30.06.2014 Views

John M. S. Bartlett.pdf - Bio-Nica.info

John M. S. Bartlett.pdf - Bio-Nica.info

John M. S. Bartlett.pdf - Bio-Nica.info

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

TM<br />

Methods in Molecular <strong>Bio</strong>logy<br />

Volume 226<br />

PCR<br />

Protocols<br />

SECOND EDITION<br />

Edited by<br />

<strong>John</strong> M. S. <strong>Bartlett</strong><br />

David Stirling


Contents<br />

1. A Short History of the Polymerase Chain Reaction 3<br />

<strong>John</strong> M. S. <strong>Bartlett</strong> and David Stirling<br />

2. PCR Patent Issues 7<br />

Peter Carroll and David Casimir<br />

3. Equipping and Establishing a PCR Laboratory 15<br />

Susan McDonagh<br />

4. Quality Control in PCR 20<br />

David Stirling<br />

5. Extraction of Nucleic Acid Templates 27<br />

<strong>John</strong> M. S. <strong>Bartlett</strong><br />

6. Extraction of DNA from Whole Blood 29<br />

<strong>John</strong> M. S. <strong>Bartlett</strong> and Anne White<br />

7. DNA Extraction from Tissue 33<br />

Helen Pearson and David Stirling<br />

8. Extraction of DNA from Microdissected Archival Tissues 35<br />

James J. Going<br />

9. RNA Extraction from Blood 43<br />

Helen Pearson<br />

10. RNA Extraction from Frozen Tissue 45<br />

<strong>John</strong> M. S. <strong>Bartlett</strong><br />

11. RNA Extraction from Tissue Sections 47<br />

Helen Pearson<br />

12. Dual DNA/RNA Extraction 49<br />

David Stirling and <strong>John</strong> M. S. <strong>Bartlett</strong><br />

13. DNA Extraction from Fungi, Yeast, and Bacteria 53<br />

David Stirling<br />

14. Isolation of RNA Viruses from <strong>Bio</strong>logical Materials 55<br />

Susan McDonagh<br />

15. Extraction of Ancient DNA 57<br />

Wera M. Schmerer<br />

16. DNA Extraction from Plasma and Serum 63<br />

David Stirling<br />

17. Technical Notes for the Detection of Nucleic Acids 65<br />

<strong>John</strong> M. S. <strong>Bartlett</strong><br />

18. Technical Notes for the Recovery and Purificationof PCR Products from Acrylamide Gels 77<br />

David Stirling<br />

19. PCR Primer Design 81<br />

David L. Hyndman and Masato Mitsuhashi<br />

20. Optimization of Polymerase Chain Reactions 89<br />

Haiying Grunenwald<br />

21. Subcycling PCR for Long-Distance Amplificationsof Regions with High and Low<br />

Guanine–Cystine Content---- Amplification of the Intron 22 Inversion of the FVIII Gene<br />

David Stirling 101<br />

22. Rapid Amplification of cDNA Ends<br />

i


Xin Wang and W. Scott Young III 105<br />

23. Randomly Amplified Polymorphic DNA Fingerprinting--The Basics 117<br />

Ranil S. Dassanayake and Lakshman P. Samaranayake<br />

24. Microsphere-Based Single NucleotidePolymorphism Genotyping 123<br />

Marie A. Iannone, J. David Taylor, Jingwen Chen, May-Sung Li,Fei Ye, and Michael P. Weiner<br />

25. Ligase Chain Reaction 135<br />

William H. Benjamin, Jr., Kim R. Smith, and Ken B. Waites<br />

26. Nested RT-PCR in a Single Closed Tube 151<br />

Antonio Olmos, Olga Esteban, Edson Bertolini, and Mariano Cambra<br />

27. Direct PCR from Serum 161<br />

kenji Abe<br />

28. Long PCR Amplification of Large Fragmentsof Viral Genomes 167<br />

---A Technical Overview<br />

Raymond Tellier, Jens Bukh, Suzanne U. Emerson, and Robert H. Purcell<br />

29. Long PCR Methodology 173<br />

Raymond Tellier, Jens Bukh, Suzanne U. Emerson, and Robert H. Purcell<br />

30. Qualitative and Quantitative PCR-----A Technical Overview 181<br />

David Stirling<br />

31. Ultrasensitive PCR Detection of Tumor Cells in Myeloma 185<br />

Friedrich W. Cremer and Marion Moos<br />

32. Ultrasensitive Quantitative PCR to Detect RNA Viruses 197<br />

Susan McDonagh<br />

33. Quantitative PCR for cAMP RI Alpha mRNA 205<br />

---- Use of Site-Directed Mutation and PCR Mimics<br />

<strong>John</strong> M. S. <strong>Bartlett</strong><br />

34. Quantitation of Multiple RNA Species 211<br />

Ron Kerr<br />

35. Differential Display----- A Technical Overview 217<br />

<strong>John</strong> M. S. <strong>Bartlett</strong><br />

36. AU-Differential Display, Reproducibilityof a Differential mRNA Display Targeted to AU Motifs 225<br />

Orlando Dominguez, Lidia Sabater, Yaqoub Ashhab,Eva Belloso, and Ricardo Pujol-Borrell<br />

37. PCR Fluorescence Differential Display 237<br />

Kostya Khalturin, Sergej Kuznetsov, and Thomas C. G. Bosch<br />

38. Microarray Analysis Using RNA Arbitrarily Primed PCR 245<br />

Steven Ringquist, Gaelle Rondeau, Rosa-Ana Risques,Takuya Higashiyama, Yi-<br />

Peng Wang, Steffen Porwollik,David Boyle, Michael McClelland, and <strong>John</strong> Welsh<br />

39. Oligonucleotide Arrays for Genotyping 255<br />

--- Enzymatic Methods for Typing Single Nucleotide Polymorphisms and Short<br />

Tandem Repeats<br />

Stephen Case-Green, Clare Pritchard, and Edwin Southern<br />

40. Serial Analysis of Gene Expression 271<br />

Karin A. Oien<br />

41. Mutation and Polymorphism Detection----- A Technical Overview 287<br />

Joanne Edwards and <strong>John</strong> M. S. <strong>Bartlett</strong><br />

42. Combining Multiplex and Touchdown PCRfor Microsatellite Analysis 295<br />

Kanokporn Rithidech and <strong>John</strong> J. Dunn<br />

ii


43. Detection of Microsatellite Instability and Lossof Heterozygosity Using DNA<br />

Extracted fromFormalin-Fixed Paraffin-Embedded Tumor Materialby Fluorescence-<br />

Based Multiplex Microsatellite PCR 301<br />

Joanne Edwards and <strong>John</strong> M. S. <strong>Bartlett</strong><br />

44. Reduction of Shadow Band Synthesis DuringPCR Amplification of Repetitive<br />

Sequencesfrom Modern and Ancient DNA 309<br />

Wera M. Schmerer<br />

45. Degenerate Oligonucleotide-Primed PCR 315<br />

Michaela Aubele and Jan Smida<br />

46. Mutation Detection Using RT-PCR-RFLP<br />

Hitoshi Nakashima, Mitsuteru Akahoshi, and Yosuke Tanaka 319<br />

47. Multiplex Amplification RefractoryMutation System for the Detectionof<br />

Prothrombotic Polymorphisms 323<br />

David Stirling<br />

48. PCR-SSCP Analysis of Polymorphism 328<br />

--- A Simple and Sensitive Method for Detecting DifferencesBetween Short Segments of DNA<br />

Mei Han and Mary Ann Robinson<br />

49. Sequencing---- A Technical Overview 337<br />

David Stirling<br />

50. Preparation and Direct Automated Cycle Sequencingof PCR Products 341<br />

Susan E. Daniels<br />

51. Nonradioactive PCR Sequencing Using Digoxigenin 347<br />

Siegfried Kösel, Christoph B. Lücking, Rupert Egensperger,and Manuel B. Graeber<br />

52. Direct Sequencing by Thermal Asymmetric PCR 355<br />

Georges-Raoul Mazars and Charles Theillet<br />

53. Analysis of Nucleotide Sequence Variationsby Solid-Phase Minisequencing 361<br />

Anu Suomalainen and Ann-Christine Syvänen<br />

54. Direct Sequencing with Highly Degenerateand Inosine-Containing Primers 367<br />

Zhiyuan Shen, Jingmei Liu, Robert L. Wells, and Mortimer M. Elkind<br />

55. Determination of Unknown Genomic SequencesWithout Cloning 373<br />

Jean-Pierre Quivy and Peter B. Becker<br />

56. Cloning PCR Products for Sequencing in M13 Vectors 385<br />

David Walsh<br />

57. DNA Rescue by the Vectorette Method 393<br />

Marcia A. McAleer, Alison J. Coffey, and Ian Dunham<br />

58. Technical Notes for Sequencing Difficult Templates 401<br />

David Stirling<br />

59. PCR-Based Detection of Nucleic Acidsin Chromosomes, Cells, and Tissues 405<br />

Technical Considerations on PRINS and In Situ PCR and Comparison with In Situ Hybridization<br />

Ernst J. M. Speel, Frans C. S. Ramaekers, and Anton H. N. Hopman<br />

60. Cycling Primed In Situ Amplification 425<br />

<strong>John</strong> H. Bull and Lynn Paskins<br />

61. Direct and Indirect In Situ PCR 433<br />

Klaus Hermann Wiedorn and Torsten Goldmann<br />

62. Reverse Transcriptase In Situ PCR----New Methods in Cellular Interrogation 445<br />

Mark Gilchrist and A. Dean Befus<br />

iii


63. Primed In Situ Nucleic Acid Labeling Combined with Immunocytochemistry to<br />

Simultaneously Localize DNA and Proteins in Cells and Chromosomes 453<br />

Ernst J. M. Speel, Frans C. S. Ramaekers, and Anton H. N. Hopman<br />

64. Cloning and Mutagenesis--- A Technical Overview 467<br />

Helen Pearson and David Stirling<br />

65. Using T4 DNA Polymeraseto Generate Clonable PCR Products 469<br />

Kai Wang<br />

66. A T-Linker Strategy for Modificationand Directional Cloning of PCR Products 475<br />

Robert M. Horton, Raghavanpillai Raju, and Bianca M. Conti-Fine<br />

67. Cloning Gene Family Members Using PCRwith Degenerate Oligonucleotide Primers<br />

Gregory M. Preston 485<br />

68. cDNA Libraries from a Low Amount of Cells 499<br />

Philippe Ravassard, Christine Icard-Liepkalns, Jacques Mallet,and Jean Baptiste<br />

Dumas Milne Edwards<br />

69. Creation of Chimeric Junctions, Deletions, and Insertions by PCR 511<br />

Genevieve Pont-Kingdon<br />

70. Recombination and Site-Directed MutagenesisUsing Recombination PCR 517<br />

Douglas H. Jones and Stanley C. Winistorfer<br />

71. Megaprimer PCR---- Application in Mutagenesis and Gene Fusion 525<br />

Emily Burke and Sailen Barik<br />

iv


History of PCR 3<br />

1<br />

A Short History of the Polymerase Chain Reaction<br />

<strong>John</strong> M. S. <strong>Bartlett</strong> and David Stirling<br />

The development of the polymerase chain reaction (PCR) has often been likened<br />

to the development of the Internet, and although this does risk overstating the impact<br />

of PCR outside the scientific community, the comparison works well on a number<br />

of levels. Both inventions have emerged in the last 20 years to the point where it is<br />

difficult to imagine life without them. Both have grown far beyond the confines of<br />

their original simple design and have created opportunities unimaginable before their<br />

invention. Both have also spawned a whole new vocabulary and professionals literate<br />

in that vocabulary. It is hard to believe that the technique that formed the cornerstone of<br />

the human genome project and is fundamental to many molecular biology laboratory<br />

protocols was discovered only 20 years ago. For many, the history and some of the<br />

enduring controversies are unknown yet, as with the discovery of the structure of DNA<br />

in the 1950s, the discovery of PCR is the subject of claim and counterclaim that has<br />

yet to be fully resolved. The key stages are reviewed here in brief for those for whom<br />

both the history and application of science holds interest.<br />

The origins of PCR as we know it today sprang from key research performed in<br />

the early 1980s at Cetus Corporation in California. The story is that in the spring of<br />

1983, Kary Mullis had the original idea for PCR while cruising in a Honda Civic on<br />

Highway 128 from San Francisco to Mendocino. This idea claimed to be the origin<br />

of the modern PCR technique used around the world today that forms the foundation<br />

of the key PCR patents. The results for Mullis were no less satisfying; after an initial<br />

$10,000 bonus from Cetus Corporation, he was awarded the 1993 Nobel Prize for<br />

chemistry.<br />

The original concept for PCR, like many good ideas, was an amalgamation of<br />

several components that were already in existence: The synthesis of short lengths of<br />

single-stranded DNA (oligonucleotides) and the use of these to direct the target-specific<br />

synthesis of new DNA copies using DNA polymerases were already standard tools in<br />

the repertoire of the molecular biologists of the time. The novelty in Mullis’s concept<br />

was using the juxtaposition of two oligonucleotides, complementary to opposite strands<br />

of the DNA, to specifically amplify the region between them and to achieve this in a<br />

repetitive manner so that the product of one round of polymerase activity was added<br />

to the pool of template for the next round, hence the chain reaction. In his History of<br />

PCR (1), Paul Rabinow quotes Mullis as saying:<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

3


4 <strong>Bartlett</strong> and Stirling<br />

The thing that was the “Aha!” the “Eureka!” thing about PCR wasn’t just putting those<br />

[things] together…the remarkable part is that you will pull out a little piece of DNA from<br />

its context, and that’s what you will get amplified. That was the thing that said, “you could<br />

use this to isolate a fragment of DNA from a complex piece of DNA, from its context.”<br />

That was what I think of as the genius thing.…In a sense, I put together elements that<br />

were already there.…You can’t make up new elements, usually. The new element, if any,<br />

it was the combination, the way they were used.…The fact that I would do it over and over<br />

again, and the fact that I would do it in just the way I did, that made it an invention…the<br />

legal wording is “presents an unanticipated solution to a long-standing problem,” that’s<br />

an invention and that was clearly PCR.<br />

In fact, although Mullis is widely credited with the original invention of PCR,<br />

the successful application of PCR as we know it today required considerable further<br />

development by his colleagues at Cetus Corp, including colleagues in Henry Erlich’s<br />

lab (2–4), and the timely isolation of a thermostable polymerase enzyme from a<br />

thermophilic bacterium isolated from thermal springs. Furthermore, challenges to the<br />

PCR patents held by Hoffman La Roche have claimed at least one incidence of “prior<br />

art,” that is, that the original invention of PCR was known before Mullis’s work in the<br />

mid-1980s. This challenge is based on early studies by Khorana et al. in the late 1960s<br />

and early 1970s (see chapter 2). Khorana’s work used a method that he termed repair<br />

replication, and its similarity to PCR can be seen in the following steps: (1) annealing<br />

of primers to templates and template extension; (2) separation of the newly synthesized<br />

strand from the template; and (3) re-annealing of the primer and repetition of the cycle.<br />

Readers are referred to an extensive web-based literature on the patent challenges<br />

arising from this “prior art” and to chapter 2 herein for further details. Whatever the<br />

final outcome, it is clear that much of the work that has made PCR such a widely<br />

used methodology arose from the laboratories of Mullis and Erlich at Cetus in the<br />

mid-1980s.<br />

The DNA polymerase originally used for the PCR was extracted from the bacterium<br />

Escherichia coli. Although this enzyme had been a valuable tool for a wide range of<br />

applications and had allowed the explosion in DNA sequencing technologies in the<br />

preceding decade, it had distinct disadvantages in PCR. For PCR, the reaction must<br />

be heated to denature the double-stranded DNA product after each round of synthesis.<br />

Unfortunately, heating also irreversibly inactivated the E. coli DNA polymerase,<br />

and therefore fresh aliquots of enzyme had to be added by hand at the start of each<br />

cycle. What was required was a DNA polymerase that remained stable during the<br />

DNA denaturation step performed at around 95°C. The solution was found when the<br />

bacterium Thermophilus aquaticus was isolated from hot springs, where it survived<br />

and proliferated at extremely high temperatures, and yielded a DNA polymerase that<br />

was not rapidly inactivated at high temperatures. Gelfand and his associates at Cetus<br />

purified and subsequently cloned this polymerase (5,6), allowing a complete PCR<br />

amplification to be created without opening the reaction tube. Furthermore, because the<br />

enzyme was isolated from a thermophilic organism, it functioned optimally at temperature<br />

of around 72°C, allowing the DNA synthesis step to be performed at higher<br />

temperatures than was possible with the E. coli enzyme, which ensured that the<br />

template DNA strand could be copied with higher fidelity as the result of a greater<br />

stringency of primer binding, eliminating the nonspecific products that had plagued<br />

earlier attempts at PCR amplification.


History of PCR 5<br />

Fig. 1. Results of a PubMed search for articles containing the phrase “Polymerase Chain<br />

Reaction.” Graph shows number of articles listed in each year.<br />

However, even with this improvement, the PCR technique was laborious and slow,<br />

requiring manual transfer between water baths at different temperatures. The first<br />

thermocycling machine, “Mr Cycle,” which replicated the temperature changes required<br />

for the PCR reaction without the need for manual transfer, was developed by Cetus<br />

to facilitate the addition of fresh thermolabile polymerases. After the purification of<br />

Taq polymerase, Cetus and Perkin–Elmer introduced the closed DNA thermal cyclers<br />

that are widely used today (7).<br />

That PCR has become one of the most widely used tools in molecular biology is<br />

clear from Fig. 1. What is not clear from this simplistic analysis of the literature is the<br />

huge range of questions that PCR is being used to answer. Another scientist at Cetus,<br />

Stephen Scharf, is quoted as stating that<br />

…the truly astonishing thing about PCR is precisely that it wasn’t designed to solve<br />

a problem; once it existed, problems began to emerge to which it could be applied. One<br />

of PCR’s distinctive characteristics is unquestionably its extraordinary versatility. That<br />

versatility is more than its ‘applicability’ to many different situations. PCR is a tool that<br />

has the power to create new situations for its use and those required to use it.<br />

More than 3% of all PubMed citations now refer to PCR (Fig. 2). Techniques have<br />

been developed in areas as diverse as criminal forensic investigations, food science,<br />

ecological field studies, and diagnostic medicine. Just as diverse are the range of<br />

adaptations and variations on the original theme, some of which are exemplified in<br />

this volume. The enormous advances made in our understanding of the human genome<br />

(and that of many other species), would not have been possible, where it not for the<br />

remarkable simple and yet exquisitely adaptable technique which is PCR.


6 <strong>Bartlett</strong> and Stirling<br />

Fig. 2. Results of a PubMed search for articles containing the phrase “Polymerase Chain<br />

Reaction.” Graph shows number of articles listed in each year expressed as a percentage of<br />

the total PubMed citations for each year.<br />

References<br />

1. Rabinow, P. (1996) Making PCR: A Story of <strong>Bio</strong>technology. University of Chicago Press,<br />

Chicago.<br />

2. Saiki, R., Scharf, S., Faloona, F., Mullis, K., Horn, G., and Erlich, H. (1985) Enzymatic<br />

amplification of beta-globin genomic sequences and restriction site analysis for diagnosis<br />

of sickle cell anemia. Science 230, 1350–1354.<br />

3. Mullis, K., Faloona, F., Scharf, S., Saiki, R., Horn, G., and Erlich, H. (1986) Specific<br />

enzymatic amplification of DNA in vitro: The polymerase chain reaction. Cold Spring<br />

Harbor Symp. Quant. <strong>Bio</strong>l. 51, 263–273.<br />

4. Mullis, K. and Faloona, F. (1987) Specific synthesis of DNA in vitro via a polymerasecatalyzed<br />

chain reaction. Methods Enzymol. 155, 335–350.<br />

5. Saiki, R., Gelfand, D., Stoffel, S., Scharf, S., Higuchi, R., Horn, et al. (1988) Primerdirected<br />

enzymatic amplification of DNA with a thermostable DNA polymerase. Science<br />

239, 487– 491.<br />

6. Lawyer, F., Stoffer, S, Saiki, R., Chang, S., Landre, P., Abramson, R., et al. (1993) Highlevel<br />

expression, purification, and enzymatic characterization of full-length Thermus<br />

aquaticus DNA polymerase and a truncated form deficient in 5′ to 3′ exonuclease activity.<br />

PCR Methods Appl. 2, 275–287.<br />

7. http://www.si.edu/archives/ihd/videocatalog/9577.htm


PCR Patent Issues 7<br />

2<br />

PCR Patent Issues<br />

Peter Carroll and David Casimir<br />

1. Introduction<br />

The science of the so-called polymerase chain reaction (PCR) is now well known.<br />

However, the legal story associated with PCR is, for the most part, not understood and<br />

constantly changing. This presents difficulties for scientists, whether in academia or<br />

industry, who wish to practice the PCR process. This chapter summarizes the major<br />

issues related to obtaining rights to practice PCR. The complexity of the patent system<br />

is explained with a few PCR-specific examples highlighted. The chapter also provides<br />

an overview of the exemption or exception from patent infringement associated with<br />

certain bona-fide researchers and discusses the status of certain high-profile patents<br />

covering aspects of the PCR process.<br />

2. Intellectual Property Rights<br />

Various aspects of the PCR process, including the method itself, are protected by<br />

patents in the United States and around the world. As a general rule, patents give the<br />

patent owner the exclusive right to make, use, and sell the compositions or process<br />

claimed by the patent. If someone makes, uses, or sells the patented invention in<br />

a country with an issued patent, the patent owner can invoke the legal system of<br />

that country to stop future infringing activities and possibly recover money from the<br />

infringer.<br />

A patent owner has the right to allow, disallow, or set the terms under which<br />

people make, use, and sell the invention claimed in their patents. In an extreme<br />

situation, a patent owner can exclude everyone from making, using, and selling the<br />

invention, even under conditions where the patent owner does not produce the product<br />

themselves—effectively removing the invention from the public for the lifetime of the<br />

patent (typically 20 years from the filing date of the patent). If a patent owner chooses<br />

to allow others to make, use, or sell the invention, they can contractually control nearly<br />

every aspect of how the invention is disbursed to the public or to certain companies<br />

or individuals, so long as they are not unfairly controlling products not covered by<br />

the patent. For example, a patent owner can select or exclude certain fields of use for<br />

methods like PCR (e.g., research use, clinical use, etc.) while allowing others.<br />

There are an extraordinary number of patents related to the PCR technology. For<br />

example, in the United States alone, there are more than 600 patents claiming aspects<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

7


8 Carroll and Casimir<br />

of PCR. Such patents cover the basic methods itself (originally owned by Cetus<br />

Corporation and now owned by Hoffmann-LaRoche), thermostable polymerases<br />

useful in PCR, as well as many non-PCR applications, (e.g., Taq polymerase, Tth<br />

polymerase, Pfu polymerase, KOD polymerase, Tne polymerase, Tma polymerase,<br />

modified polymerases, etc.), devices used in PCR (e.g., thermocyclers, sample tubes<br />

and vessels, solid supports, etc.), reagents (e.g., analyte-specific amplification primers,<br />

buffers, internal standards, etc.), and applications involving the PCR process (e.g.,<br />

reverse-transcription PCR, nested PCR, multiplex PCR, nucleic acid sequencing, and<br />

detection of specific analytes). This collection of patents is owned by a wide variety<br />

of entities, including government agencies, corporations, individual inventors, and<br />

universities. However, the most significant patents (see Table 1), covering the basic<br />

PCR method, the most widely used polymerase (Taq polymerase), and thermocyclers,<br />

are assigned to Hoffmann-LaRoche and are controlled by Hoffmann-LaRoche or<br />

Applera Corporation (previously known as PE/Applied <strong>Bio</strong>systems) and are available<br />

to the public through an intricate web of licenses.<br />

3. Navigating the PCR Patent Minefield<br />

The following discussion focuses on issues related to the earliest and most basic<br />

PCR-related patents. A full analysis of the hundreds of PCR-related patents is not<br />

practical in an article this size, let alone a multivolume treatise. It is hoped that the<br />

following discussion will provide a preliminary framework for understanding the broad<br />

PCR patent landscape.<br />

The early PCR patents now owned by Hoffmann-LaRoche have been aggressively<br />

enforced. In particular, the earliest patents intended to cover the basic PCR method and<br />

the Taq polymerase enzyme (U.S. Patent No. 4,683,202 to Kary Mullis, U.S. Patent<br />

No. 4,683,195 to Kary Mullis et al., U.S. Patent No. 4,889,818 to Gelfand et al. and<br />

foreign counterparts) have regularly been litigated and used to threaten litigation,<br />

even against academic researchers. This aggressive patent stance has created an<br />

environment of fear, confusion, and debate, particularly at universities and among<br />

academic researchers. Because of this aggressive patent enforcement, issues with<br />

respect to these patents are most relevant and are focused on herein.<br />

3.1. Obtaining Rights to Practice PCR<br />

In the case of the early PCR patents, Hoffmann-LaRoche, directly and through<br />

certain designated partners, has made PCR available to the public under specific conditions,<br />

depending on the intended use of the method (see for availability of licenses and current details). For example, for<br />

nonsequencing research use, PCR users have two options. They can individually<br />

negotiate a license from Applera (a proposition that is impractical for many researchers).<br />

Optionally, they can purchase “certain reagents” from a “licensed supplier” in<br />

conjunction with the use of “an authorized thermal cycler.” This essentially means<br />

that the user must purchase thermostable enzymes and thermocyclers from suppliers<br />

licensed by Hoffmann-LaRoche or Applera. Not surprisingly, the price of these<br />

products from licensed suppliers greatly exceeds the price of equivalent products from<br />

nonlicensed suppliers. Indeed, thermostable enzymes from licensed suppliers may


PCR Patent Issues 9<br />

Table 1<br />

PCR Patents<br />

U.S. patent Issue Expiration Related international<br />

number date date patents Claimed technology<br />

4,683,195 07/28/87 03/28/05 Australia: 591104B Amplification methods<br />

Australia: 586233B<br />

Canada: 134012B<br />

Europe: 200362B<br />

Europe: 201184B<br />

Europe: 505012B<br />

Japan: 2546576B<br />

Japan: 2622327B<br />

Japan: 4067957B<br />

Japan: 4067960B<br />

4,683,202 07/28/87 03/28/05 Same as 4,683,195 Amplification methods<br />

4,965,188 10/23/90 03/28/05 Australia: 586233B Amplification methods using<br />

Australia: 591104B<br />

thermostable polymerases<br />

Australia: 594130B<br />

Australia: 632857B<br />

Canada: 1340121B<br />

Europe: 200362B<br />

Europe: 201184B<br />

Europe: 237362B<br />

Europe: 237362B<br />

Europe: 258017B<br />

Europe: 459532B<br />

Europe: 505012B<br />

Japan: 2502041B<br />

Japan: many others<br />

4,889,818 12/26/86 12/26/06 Australia: 632857B Purified Taq polymerase<br />

(currently Europe: 258017B enzyme<br />

unenforceable) Japan: 2502041B<br />

Japan: 2502042B<br />

Japan: 2719529B<br />

Japan: 3031434B<br />

Japan: 5074345B<br />

Japan: 8024570B<br />

5,079,352 01/07/92 01/07/09 Same as 4,889,818, Recombinant Taq<br />

plus<br />

polymerase enzyme<br />

Europe: 395736B<br />

and fragments<br />

Japan: 2511548B<br />

Japan: 2511548B<br />

5,038,852 8/13/91 08/13/08 Australia: 612316B Apparatus and method for<br />

Australia: 653932B<br />

performing automated<br />

Europe: 236069B<br />

amplification<br />

Japan: 2613877B


10 Carroll and Casimir<br />

cost more than twice as much as from nonlicensed suppliers (1). This elevated cost<br />

can place a substantial financial burden on researchers who require heavy PCR usage,<br />

particularly academic researchers on fixed and limited grant budgets. To the extent<br />

universities require their researchers to used licensed products, the aggregate cost<br />

increase for many large research universities is substantial. (For a list of Taq polymerase<br />

suppliers and prices, including licensed and unlicensed suppliers, see Constans, ref. 2).<br />

3.2. Bona-Fide Researchers Are Not Infringers<br />

As mentioned previously, Hoffmann-LaRoche has taken the position that academic<br />

researchers are infringers of their patents if they are not meeting the prescribed licensing<br />

requirements (e.g., not purchasing authorized reagents and equipment). At one point.<br />

several years ago, Hoffmann-LaRoche specifically named more than 40 American<br />

universities and government laboratories and more than 200 individual scientists as<br />

directly infringing certain patents through their basic research (3). Voicing the view<br />

of many researchers, Dr. Arthur Kornberg, professor emeritus at Stanford University<br />

and Nobel laureate, has stated that the actions by Hoffmann-LaRoche to restrain the<br />

use and extension of PCR technology by universities and nonprofit basic research<br />

institutions “violated practices and principles basic to the advancement of knowledge<br />

for the public welfare.”<br />

Fortunately for academic researchers, the laws of the United States and other<br />

jurisdictions agree with Dr. Kornberg. US patent law recognizes an exemption or<br />

exception from infringement associated with bona-fide research (i.e., not-for-profit<br />

activities). The experimental use exception to the patent infringement provisions of<br />

US law has its origins in the notion that “it could never have been the intention of<br />

the legislature to punish a man, who constructed…a [patented] machine merely for<br />

philosophical experiments….” (4). An authoritative discussion on the research use<br />

exception appears in the case Roche Prods., Inc. v Bolar Pharmaceutical Co. (5).<br />

Even though this case is generally considered to restrict the scope of the research use<br />

exemption (failing to find noninfringement where the defendant’s acts were “solely<br />

for business reasons”), the case makes it clear that the exception is alive and well<br />

where the acts are “for amusement, to satisfy idle curiosity, or for strictly philosophical<br />

inquiry.” Thus, to the extent that researchers’ use of PCR is not applied to commercial<br />

applications or development (e.g., for-sale product development, for-profit diagnostic<br />

testing), the researchers cannot be considered infringers. For example, pure basic<br />

research, which describes most university research, cannot be considered commercial,<br />

and the researchers are not infringers. This applied to the PCR patents, as well as any<br />

other patent. Hoffmann-LaRoche has taken the position that “These researchers…are<br />

manifestly in the business of doing research in order to…attract private and government<br />

funding through the publication of their experiments in the scientific literature, create<br />

patentable inventions, and generate royalty income for themselves and their institutions<br />

through the licensing of such invention.” However, current US law does not support<br />

this extraordinarily broad view of commercial activity, and Hoffmann-LaRoche seems<br />

to be alone in making such broad assertions.<br />

Although the above discussion relates to the United States, researchers in other<br />

countries may or may not have the same exemption. The scope of this article does not<br />

permit a country-by-country analysis. However, it must be noted that many countries


PCR Patent Issues 11<br />

are in alignment with the position taken by US courts or provide an even broader<br />

exemption. For example, it is not considered an infringement in Canada to construct a<br />

patented article for the purpose of improving upon it or to ascertain whether a certain<br />

addition, subtraction, or improvement on it is workable. The Supreme Court of Canada<br />

spoke on this issue stating that “[N]o doubt if a man makes things merely by way<br />

of bona fide experiment, and not with the intention of selling and making use of the<br />

thing so made for the purpose of which a patent has been granted, but with the view of<br />

improving upon the invention the subject of the patent, or with the view of assessing<br />

whether an improvement can be made or not, that is not an invasion of the exclusive<br />

rights granted by the patent. Patent rights were never granted to prevent persons<br />

of ingenuity exercising their talents in a fair way.” Likewise, UK law provides an<br />

exemption from infringement for acts that are performed privately and for purposes<br />

that are not commercial and for acts performed for experimental purposes relating to<br />

the subject matter of the invention. The experimental purposes may have a commercial<br />

end in view, but they are only exempt from infringement if they relate to the subject<br />

matter of the invention. For example, it has been held by the UK courts that trials<br />

conducted to discover something unknown or to test a hypothesis, to find out whether<br />

something which is known to work in specific conditions would work in different<br />

conditions, or even perhaps to see whether the experimenter could manufacture commercially<br />

in accordance with the patent can be regarded as experiments and exempted<br />

from infringement. Researchers in any particular country who wish to obtain current<br />

<strong>info</strong>rmation about their ability to conduct research projects without incurring patent<br />

infringement liability should contact the patent office or an attorney in their respective<br />

countries. Unfortunately, there is very little literature addressing these issues, and<br />

because the law is constantly changing, older articles may not provide accurate<br />

<strong>info</strong>rmation.<br />

Even with uncertainties, it is clear that in many locations, researchers conducting<br />

basic research without a commercial end are free to practice in their field without<br />

fear or concern about the patent rights of others. Researchers at corporations likely<br />

cannot take advantage of such an infringement exemption. For researchers involved<br />

in work with a commercial link (e.g., researchers at private corporations, diagnostic<br />

laboratories reporting patient results for fees, academic research laboratories with<br />

private corporate collaborations, and the like), a license may be required. Unfortunately,<br />

each case needs to be evaluated on its own facts to determine whether a license is<br />

required and no general formula can be given. However, many corporations have<br />

personnel responsible for analyzing the need for, and acquisition of, patent rights. As<br />

such, bench scientists can generally go about their work without the burden of worrying<br />

about patent rights, or at a minimum, need only know the basic principles and issues so<br />

as to <strong>info</strong>rm the appropriate personnel if potential patent issues arise.<br />

3.3. Not Every Patent Is a Valid Patent<br />

In addition to the experimental use exception, researchers, including commercial<br />

researchers, may obtain freedom from the early PCR patents because of problems with<br />

the patents themselves. Although issued patents are presumed valid and are enforceable<br />

until a court of law says otherwise, the early PCR patents have begun to fall under<br />

scrutiny and may not be upheld in the future such that the basic reagents and methods


12 Carroll and Casimir<br />

are no longer covered by patents. It must be emphasized that at this time most of<br />

the patents are still deemed valid and enforceable. However, researchers may wish<br />

to follow the events as they unfold with respect to the enforceability and validity of<br />

the PCR patents.<br />

The first blow against the PCR patents was struck by Promega Corporation (Promega;<br />

Promega Corporation is a corporation headquartered in Madison, Wisconsin that<br />

produces for sale reagents and other products for the life science community.).<br />

HoffmannLaRoche filed an action against Promega on October 27, 1992 alleging<br />

breach of a contract for the sale of Taq DNA Polymerase (Taq), infringement of<br />

certain patents—the PCR Patents (United States Patent Nos. 4,683,195 and 4,683,202)<br />

and United States Patent No. 4,889,818—and related causes of action. At issue was<br />

United States Patent No. 4,889,818 (the ‘818 patent), entitled “Purified Thermostable<br />

Enzyme.” Promega denied the allegations of the complaint and claimed, among other<br />

things, that the ‘818 patent was obtained by fraud and was therefore unenforceable.<br />

After a trial in 1999, a US court held that all of the claims of the ‘818 patent unenforceable<br />

for inequitable conduct or fraud. The unenforceable claims are provided below.<br />

1. Purified thermostable Thermus aquaticus DNA polymerase that migrates on a denaturing<br />

polyacrylamide gel faster than phosphorylase B and more slowly than does bovine serum<br />

albumin and has an estimated molecular weight of 86,000 to 90,000 Dalton when compared<br />

with a phosphorylase B standard assigned a molecular weight of 92,500 Dalton.<br />

2. The polymerase of claim 1 that is isolated from Thermus aquaticus.<br />

3. The polymerase of claim 1 that is isolated from a recombinant organism transformed with<br />

a vector that codes for the expression of Thermis aquaticus DNA polymerase.<br />

The court concluded that Promega had demonstrated by clear and convincing<br />

evidence that the applicants committed inequitable conduct by, among other things,<br />

withholding material <strong>info</strong>rmation from the patent office; making misleading statements;<br />

making false claims; misrepresenting that experiments had been conducted when, in<br />

fact, they had not; and making deceptive, scientifically unwarranted comparisons. The<br />

court concluded that those misstatements or omissions were intentionally made to<br />

mislead the Patent Office. The court’s decision has been appealed, and a decision from<br />

the Federal Circuit Court of Appeals is expected shortly. Pending the appeal court<br />

decision, the ‘818 patent is unenforceable.<br />

Patents have also been invalidated in Australia and Europe. On November 12,<br />

1997, the Australian Patent Office invalidated all claims concerning native Taq DNA<br />

polymerase and DNA polymerases from any other Thermus species, contained in a<br />

patent held by Hoffmann-La Roche (application no. 632857). The Australian Patent<br />

Office concluded that the enzyme had been previously purified in Moscow and<br />

published by Kaledin et al. (6) and that certain patent claims were unfairly broad.<br />

Although the case has been appealed, as of this writing, the Taq patent in Australia<br />

is unenforceable.<br />

In Europe, on May 30, 2001, the opposition division of the European Patent Office<br />

held that claims in the thermostable enzyme patent EP 0258017B1 (a patent equivalent<br />

to the ‘818 patent in the United States) were unpatentable because they lacked an<br />

inventive step in view of previous publications to Kaledin et al. (6) and Chient et al.<br />

(7), as well as knowledge generally know in the field at the time the patent application<br />

was filed.


PCR Patent Issues 13<br />

Although it has not been determined yet whether the PCR method patents were<br />

procured with the same types of misleading and deceptive behavior, the PCR patents<br />

have been challenged based on an earlier invention by Dr. Gobind Khorana and<br />

coworkers in the late 1960s and early 1970s. Under US and many international patent<br />

laws, patent claims are not valid if they describe an invention that was used and/or<br />

disclosed by others prior to the filing date of the patent. The principle behind such rules<br />

is to prevent people from patenting, and thus removing from the public domain, things<br />

that the public already owns. Although the PCR patents make no mention of such work,<br />

DNA amplification and cycling reactions were conducted many years before the filing<br />

of the PCR patents in the laboratory of Dr. Khorana. Dr. Khorana’s method, which he<br />

called “repair replication,” involved the steps of the following: (1) extension from a<br />

primer annealed to a template; (2) separating strands; and (3) reannealing of primers to<br />

template to repeat the cycle. Dr. Khorana did not patent this work. Instead he dedicated<br />

it to the public. Unfortunately, at the time that Dr. Khorana discovered his amplification<br />

process, it was not practical to use the method for nucleic acid amplification, and the<br />

technique did not take off as a commercial method. At the time this work was disclosed,<br />

chemically synthesized DNA for use as primers was extremely expensive and costprohibitive<br />

for even limited use. Additionally, recombinantly produced enzymes were<br />

not available. Thus, not until the 1980s, when enzyme and oligonucleotide production<br />

became more routine, could one economically replicate Dr. Khorana’s methods.<br />

The validity of the PCR patents was challenged in 1990 by E.I. Dupont De Nemours<br />

& Co. (Dupont). Based on publicly available records, it appears that Dupont pointed to<br />

the work from Dr. Khorana’s laboratory, arguing that all of the method steps required<br />

in the basic PCR method were taught by Dr. Khorana’s publications and were in fact in<br />

the public domain. Hoffmann-LaRoche (who was positioned to acquire the technology)<br />

out-maneuvered Dupont by putting the Khorana papers in front of the United States<br />

Patent and Trademark Office in a reexamination procedure. Under reexamination, the<br />

patent holder has the ability to argue the patentability of an invention to the patent<br />

office without any input allowed by third parties, such as Dupont. As shown by<br />

publicly available records, during the reexamination procedure expert declarations<br />

were entered to raise doubt about the teaching of the Khorana references. As a result<br />

(not surprisingly), the Patent Office upheld the patents. Once a patent has issued<br />

in view of a reference, there is a strong presumption of validity that courts must<br />

acknowledge in any proceedings that later attempt to invalidate the patent in view<br />

of the reference.<br />

In addition to the disadvantage caused by the reexamination procedure, publicly<br />

available records show that Dupont was not able to use several pieces of compelling<br />

evidence against the PCR patents. Dupont, although performing clever replication work<br />

to show the sufficiency of Dr. Khorana’s disclosures (in direct contrast to the expert<br />

declarations submitted to the Patent Office during reexamination), did not submit<br />

the data in a timely manner in the proceedings. The judge ruled that the data should<br />

be excluded as untimely and prejudicial. Dupont also found additional references<br />

disclosing the earlier invention by Khorana, but did not provide them to the court in<br />

time and they were not considered. Thus, it seems that validity of the PCR patents<br />

was never truly tested in view of the work conducted by Dr. Khorana and his colleagues.<br />

Such a test, as well as others, may come in the near future as part of the<br />

Promega/HoffmannLaRoche litigation.


14 Carroll and Casimir<br />

Should these or any additional patents be found invalid and unenforceable, the patent<br />

issues for researchers wishing to practice PCR will be greatly simplified. Interestingly,<br />

if it is found that one or more of the invalid or unenforceable patents were used to<br />

suppress competition in the market or to unfairly control the freedom of researchers,<br />

companies exerting such unfair market control may be subject to laws designed to<br />

prevent unfair and anticompetitive behavior. If a court were to rule that anticompetitive<br />

behavior was exercised, the violating patent owner may be forced to compensate those<br />

that were harmed. Although it is impossible to predict at this time the outcome of future<br />

court proceedings, researchers may wish to follow the progress of these cases. At a<br />

minimum, they offer perspective into the patent world and provide important subject<br />

matter for debate that is extremely relevant to shaping the future of patent public policy,<br />

an area that will increasingly play a role in the day-to-day lives of scientists.<br />

References<br />

1. Beck, S. (1998) Do you have a license? Products licensed for PCR in research applications.<br />

The Scientist 12, 21.<br />

2. Constans, J. (2001) Courts cast clouds over PCR pricing. The Scientist 15, 1.<br />

3. Finn, R. (1996) Ongoing patent dispute may have ramifications for academic researchers.<br />

The Scientist 10, 1.<br />

4. Wittemore v Cutter, 29 F. Cas. 1120 (C.C.D. Mass. 1813)(No. 17,600)(Story, J.).<br />

5. Roche Prods., Inc. v Bolar Pharmaceutical Co., 733 F.2d 858 (Fed. Cir. 1984).<br />

6. Kaledin, A. S., Sliusarenko, A. G., and Gorodetskii, S. I. (1980) Isolation and properties of<br />

DNA polymerase from extreme thermophylic bacteria Thermus aquaticus YT-1. <strong>Bio</strong>khimiia<br />

45, 644–651. In Russian.<br />

7. Chien, A., Edgar, D. B., and Trela, J. M. (1976) Deoxyribonucleic acid polymerase from the<br />

extreme thermophile Thermus Aquaticus. J. Bacteriol. 127, 1550–1557.


Equipping, Establishing a PCR Laboratory 15<br />

3<br />

Equipping and Establishing a PCR Laboratory<br />

Susan McDonagh<br />

1. Introduction<br />

Polymerase chain reaction (PCR) is a very sensitive method of amplifying specific<br />

nucleic acid, but the system is susceptible to contamination from extraneous or<br />

previously amplified DNA strands (1,2). Many specific copies of DNA are produced<br />

from each round of amplification (3) with a single aerosol containing up to 24,000<br />

copies of amplified material (4). The most important consideration when designing<br />

and equipping a laboratory for PCR is therefore to minimize the risk of contamination<br />

and ensure accurate results (5,6). To do this, it is necessary to physically separate the<br />

different parts of the process and arrange them in a unidirectional workflow (4). This<br />

avoids back flow of traffic and, along with restricted access, will reduce the risk of<br />

contamination and inaccurate results.<br />

The way in which the workflow is arranged will depend on the amount of available<br />

space. If possible, different rooms should be used for reagent preparation, sample<br />

preparation, PCR (some also separate primary and secondary stages), and post-PCR<br />

processing (see Fig. 1). Each of these areas should contain dedicated equipment,<br />

protective clothing, and consumables (1). Disposable gloves should be readily available<br />

for frequent changing to avoid cross contamination, and control material should be<br />

included in every run to monitor any contamination problems (3).<br />

2. Equipment<br />

A list of basic equipment required for a PCR laboratory is given in Table 1.<br />

2.1. Thermocyclers<br />

This is obviously the most important piece of equipment in the laboratory, with<br />

many products available from different manufacturers. Thermocyclers can be supplied<br />

with a variety of reaction vessel formats, including 0.2- and 0.5-mL microtubes; strips<br />

of tubes; microtiter plates containing up to 384 wells; glass slides; and capillaries.<br />

Temperature ramp rates and uniform heat distribution across the block are important<br />

for consistent performance. These options, along with the consideration of laboratory<br />

requirements, are factors when purchasing a machine, and these specifications are obviously<br />

reflected in the cost. For example, if basic PCR is all that is required, equipment<br />

from the lower end of the range might suffice. These machines have programmable<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

15


16 McDonagh<br />

Fig. 1. Unidirectional flow in a PCR laboratory.<br />

blocks, often with a heated lid, and a basic repertoire of cycling capabilities. If high<br />

throughput using many different protocols is required in a diagnostic setting, then a<br />

multiblock system with the advantage of adding satellite units may be appropriate.<br />

More specialized machines with gradient blocks suitable for rapid optimization studies<br />

or with specialized blocks for in situ PCR are also available.<br />

Advances in technology have resulted in the development of real-time PCR systems,<br />

which allow rapid cycling (50 cycles in less than 30 min). These systems are expensive<br />

but provide benefits, including rapid throughput, efficient optimization, and further<br />

reducing the risk of contamination with reactions and product analysis occurring in<br />

a single tube.<br />

2.2. Additional Equipment<br />

Dedicated equipment for each area of the laboratory can be purchased from regular<br />

laboratory suppliers. Contamination can often arise from breaks and spills in equipment,<br />

such as centrifuges and waterbaths (4); therefore, important considerations include<br />

the purchase of equipment that can be easily taken apart for decontamination (see<br />

Note 1).<br />

All areas require dedicated pipets (1). Plugged tips used with traditional pipets are<br />

generally cheaper and easier to use than positive displacement pipets (see Note 2).<br />

Storage space at 4°C and –20°C should be available in each area, along with access<br />

to –70°C freezer facilities.<br />

Laminar hoods are not always recommended, except at the sample extraction stage,<br />

where they are required to protect the worker. Using individual workstations with


Equipping, Establishing a PCR Laboratory 17<br />

Table 1<br />

Equipment Required<br />

Reagent Sample 1° 2°<br />

All preparation preparation PCR PCR Post-PCR<br />

Pipets Microfuge Microfuge Cyclers Cyclers Electrophoresis tanks<br />

Refrigerator Vortex Vortex Power packs<br />

–20°C freezer dH 2 O source Laminar cabinet Microwave<br />

Work stations Ice machine Gel viewing system<br />

Balance<br />

Gel documentation system<br />

pH meter<br />

decontamination facilities reduces airflow throughout the laboratory and minimizes<br />

aerosol dispersal. These may simply consist of a disposable or wipeable tray on<br />

which the worker completes all operations before treating to remove any potential<br />

contaminating nucleic acids (1,2) (see Note 3). Some manufacturers produce purposebuilt<br />

cabinets, which incorporate several decontamination and safety features.<br />

An ice machine, distilled water supply, balance, and pH meter are required in the<br />

reagent preparation area, and a microwave is ideal for melting agarose for gel assembly<br />

in the post-PCR area.<br />

2.3. Consumables<br />

Disposable plastics rather than reusable glass should be used wherever possible,<br />

and high-quality consumables, for example, Rnase-free plasticware, should be used<br />

throughout the laboratory. It is also important to note that performance may be affected<br />

by different products from different suppliers, which was demonstrated in a study in<br />

which varying results were obtained when using microtubes supplied by a number of<br />

manufacturers (2). Other factors have an inhibitory effect on PCR performance and<br />

should also be considered. Examples include methods, such as ultraviolet irradiation,<br />

which can affect reagents such as mineral oil (7), therefore it is important to avoid<br />

exposure, and powder in gloves, which has been shown to inhibit PCR (2,8); therefore,<br />

powder-free varieties are recommended (see Note 4).<br />

3. Laboratory Layout<br />

Work within the laboratory should be confined to the specific areas identified for<br />

that part of the procedure. Each of these areas is described below, but several points<br />

apply to all. These include removal of laboratory coat and gloves before moving into<br />

another part of the laboratory; provision of gloves for frequent change; avoidance of<br />

aerosols and drips; and decontamination of working area and equipment before and<br />

after use (3) (see Note 3). All reagents necessary for each process should be stored<br />

within the area in which the work is being performed (3).<br />

3.1. Reagent Preparation Laboratory<br />

This area should be kept entirely free from samples and other potential sources of<br />

nucleic acid. Stock solutions and reagents should be made up, or diluted if purchased<br />

as concentrate, then dispensed in single use aliquots (1,3,4) or small volumes (7) and


18 McDonagh<br />

stored. This means that that they can be identified and discarded if contamination<br />

does arise (9). Master mixes are made up here and added to reaction vessels before<br />

continuing onto the next stage of the process (see Note 5). If necessary, an oil overlay<br />

can also be added at this stage.<br />

3.2. Sample Preparation Laboratory<br />

Laminar flow cabinets are necessary for dealing with samples until they are<br />

inactivated and extracted, and these and other equipment should be decontaminated<br />

before and after each procedure (see Note 3). The equipment necessary will depend on<br />

the extraction methods used, but a microfuge, heating block, and vortex are minimal<br />

requirements.<br />

3.3. PCR Laboratory<br />

Primary and secondary PCR steps should be separated, preferably in different rooms,<br />

and certainly with separate thermocyclers; however, the layout of this area will depend<br />

on space and equipment available. Primary reactions containing master mix and nucleic<br />

acid should be assembled and placed on the appropriate thermocycler. After cycling,<br />

these are removed to the secondary PCR area, where reactions are assembled and<br />

placed on cyclers dedicated for this process. Other automated/integrated/single-round<br />

equipment should be positioned with secondary thermocyclers to reduce the risk of<br />

contamination (see Note 6).<br />

3.4. Post-PCR Processing<br />

All final amplified products should be dealt with in this area, which can be used<br />

for techniques, including electrophoresis, restriction fragment length polymorphism<br />

(RFLP), hybridization work, cloning, and sequencing. It is important that nothing from<br />

this area should go back through other areas involving preliminary steps but should be<br />

processed through a waste management or discard area.<br />

4. Notes<br />

1. For example, hot blocks are easier to decontaminate on a regular basis and are therefore<br />

a better option than water baths.<br />

2. Normal tips can be used for post-PCR steps.<br />

3. An ultraviolet irradiation source is valuable in reducing contamination; however, Cimino<br />

et al. (10) recommend caution when using this method alone. Otherwise, wash down all<br />

nonmetal surfaces with 0.1 N HCl, or 10% bleach, followed by water.<br />

4. Nitrile gloves should be used for safety when handling ethidium bromide if used in gel<br />

electrophoresis.<br />

5. As kit-based formats become available, reagent and master mix will be supplied, completely<br />

reducing the need for this area.<br />

6. This setup will become more difficult as combined extraction/amplification and detection<br />

equipment become more available.<br />

References<br />

1. Wolcott, M. J. (1992) Nucleic acid-based detection methods. Clin. Microbiol. Rev. 5,<br />

370–386.<br />

2. Wilson, I. G. (1997) Inhibition and facilitation of nucleic acid amplification. Appl. Environ.<br />

Microbiol. 63, 3741–3751.


Equipping, Establishing a PCR Laboratory 19<br />

3. Sambrook, J., Fritsch, E. F., and Maniatis, T. (1989) In vitro application of DNA by the<br />

polymerase chain reaction, in Molecular Cloning: A Laboratory Manual. 2nd ed., Cold<br />

Spring Harbor Laboratory Press, New York, pp. 14.14.<br />

4. Orrego, C. (1990) Organizing a laboratory for PCR work, in PCR Protocols: A Guide to<br />

Methods and Applications (Innes, M. L., Gelfand, D. H., Sninsky, J. J., White, T. J., eds.),<br />

Academic Press Inc., San Diego, pp. 447– 454.<br />

5. Baselski, V. S. (1996) The role of molecular diagnostics in the clinical microbiology<br />

laboratory. Clin. Lab. Med. 16, 49–60.<br />

6. Lisby, G. (1999) Application of nucleic acid amplification in clinical microbiology. Mol.<br />

<strong>Bio</strong>technol. 12, 75–79.<br />

7. Hughes, M. S., Beck, L. A., and Skuse, R. A. (1994) Identification and elimination of DNA<br />

sequences in Taq DNA polymerase. J. Clin. Microbiol. 32, 2007–2008.<br />

8. De Lomas, J. G., Sunzeri, F. J., and Busch, M. P. (1992) False negative results by polymerase<br />

chain reaction due to contamination by glove powder. Transfusion 32, 83–85.<br />

9. Madej, R. and Scharf, S. (1990) Basic equipment and supplies, in PCR protocols: A Guide<br />

to Methods and Applications (Innes, M. L., Gelfand, D. H., Sninsky, J. J., White, T. J.,<br />

eds.), Academic Press Inc., San Diego, pp. 455–459.<br />

10. Cimino, G. D., Metchette, K., Isaacs, S. T., and Zhu, Y. S. (1990) More false positive<br />

problems. Nature 345, 773–774.


20 McDonagh


Quality Control 21<br />

4<br />

Quality Control in PCR<br />

David Stirling<br />

1. Introduction<br />

Polymerase chain reaction (PCR), like any laboratory procedure, can be subject<br />

to a range of experimental or procedural error. A clear consideration of where such<br />

potential errors may occur is essential to minimize their impact. Careful quality control<br />

of equipment and reagents is essential.<br />

2. Equipment<br />

The previous chapter dealt with the sort of equipment that is required to perform<br />

PCR. It is commonplace for an individual laboratory to contain many sets of equipment,<br />

each bought from different manufacturers, at different times, and subjected to various<br />

amounts of abuse from students who don’t know any better and laboratory managers<br />

who do! In an ideal world, and any diagnostic or commercial laboratory, each piece of<br />

equipment should be serviced and calibrated on a regular basis, with careful records<br />

being kept of this maintenance. Unfortunately, not every laboratory has funds for fullservice<br />

contracts on all equipment. There are a few fundamental procedures, however,<br />

which will reduce errors from equipment problems.<br />

• Be consistent in the equipment used for any given PCR. If it works on Monday but not<br />

Tuesday, this may simply be to the result of using a different PCR block. Even the most<br />

modern and expensive thermal cyclers deteriorate with age.<br />

• Check pipetting devices on a regular basis (weekly is not excessive) to ensure they pipet<br />

the correct volume. This is easily performed by pipetting and weighing water. Most<br />

manufacturers produce inexpensive service packs for their pipettors.<br />

3. Reagents<br />

As with all laboratory procedures, it generally pays dividends to use high-quality<br />

reagents from reputable suppliers. You may well know someone who brews their<br />

own Taq polymerase in a vat in the garage, but do they control for batch-to-batch<br />

variability?<br />

The design of optimum PCR primers will be discussed later. It is important to<br />

remember, however, that these are single-stranded DNA molecules and are therefore<br />

relatively labile. Repeated freeze/thawing will cause degradation to shorter products,<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

21


22 Stirling<br />

which will either not anneal, or if the annealing temperature is low enough, will anneal<br />

promiscuously, yielding multiple products. Simple aliquoting primers into manageable<br />

volumes will reduce both the scope for contamination and degradation. This practice<br />

should also be adopted for dNTP stocks for the same reason.<br />

4. Operator Errors<br />

Anyone involved in teaching molecular techniques hears the same complaint again<br />

and again: “These reaction volumes are too small! . . . I can’t see a microliter!” Even for<br />

those with many years of laboratory experience (perhaps especially for those), it can be<br />

difficult to adjust to dealing with small volume reactions. Although the obvious answer<br />

may be to increase the volume, this has both cost and efficiency implications.<br />

• Use appropriate pipetting devices. A pipettor designed for the 20- to 200-µL range will<br />

not accurately dispense 10 µL.<br />

• The use of master mixes not only reduces the dependence on accurately pipetting small<br />

volumes but also improves the control over reaction contents.<br />

• Practice with the same reaction until consistent results are obtained.<br />

5. PCR-Specific Difficulties<br />

Although much of the above could apply to any analytical laboratory technique,<br />

PCR also is subject to the confounding problem of contamination. Cross contamination<br />

of samples is of concern in any discipline, and good laboratory practice, such as<br />

careful pipetting and the constant changing of disposable pipet tips, will minimize<br />

the opportunity of this occurring. Where PCR differs from most other procedure is in<br />

the production of vast quantities of the analyte during the procedure. The presence of<br />

billions of copies of potential template can create severe problems. These problems can<br />

be minimized by physically separating the pre- and postamplification processes (ideally<br />

in different rooms with different pipettors, etc.); however, they should constantly be<br />

monitored by the inclusion of appropriate controls.<br />

6. Controls<br />

• No DNA. Although it can seem extravagant to constantly set up reactions without template,<br />

this is the best way to monitor for contamination. A separate “no DNA” control should<br />

be set up for each master mix or each individual reaction. If contamination is discovered,<br />

the pipettor should be decontaminated (as per manufacturers guidelines), and the reagent<br />

aliquot should be rechecked or discarded.<br />

• Positive control. PCR is often used simply to detect the presence of specific sequence. In<br />

such circumstances, it is essential to include at least one reaction with a template known<br />

to contain the sequence.<br />

• Internal control. Even when master mixes have been used to ensure consistency of reaction<br />

components, and a positive control is used, there is the possibility that template may be<br />

omitted from individual tubes. This can be addressed by the inclusion within each reaction<br />

tube of primers, which will amplify a target known to consistently be present in the test<br />

DNA (see factor IX in Chapter 47 for example).<br />

7. Regional Quality-Assurance Programs<br />

In addition to the in-house precautions detailed above, there are a growing number<br />

of specialist quality-assurance programs that have been developed for most diagnostic<br />

PCR.


Quality Control 23<br />

These programs distribute test material to laboratories, who then report their results<br />

centrally. Results from all participating centers are compared and confidential reports<br />

issued to each center. If such a program exists in your field, details will probably be<br />

available on the Internet: join it. If one doesn’t already exist, consider starting one; it<br />

may generate enough revenue to pay for instrument service contracts!


24 Stirling


Extraction of Nucleic Acid Templates 27<br />

5<br />

Extraction of Nucleic Acid Templates<br />

<strong>John</strong> M. S. <strong>Bartlett</strong><br />

The following series of short technical descriptions covers the extraction of DNA<br />

and RNA from various starting materials. We have gathered these together at the<br />

beginning of this book to provide an easy reference. The polymerase chain reaction<br />

(PCR) techniques described in the rest of this book are, in the main, worked laboratory<br />

methods given detailed examples of the procedures used by the authors in their own<br />

research. However, the aim is to provide the reader with a method that may be translated<br />

into their own research; thus, although the description of ultrasensitive PCR focuses<br />

on viral genomes and cancer, this method may of course be equally applied to DNA<br />

from other sources. Rather than leaving the reader who is interested in applying this<br />

technique to ancient DNA or DNA from bone, etc., to search through the various<br />

chapters to find such a technique, we have collected these together for reference here.<br />

PCR provides a simple method for the amplification and analysis of DNA; however,<br />

for most applications involving PCR, the DNA (or cDNA for RNA methods) must be<br />

in a reasonably pure state. Therefore, the first stage of any experimental procedure<br />

involving PCR based technologies is the provision of a pure suspension of nucleic<br />

acids, either RNA or DNA.<br />

Extraction of nucleic acids is a fundamental precursor to almost all the techniques<br />

described within this volume. Isolation of RNA and DNA from blood and<br />

fresh tissues can be performed using a variety of techniques, which also form the<br />

basis of methods of extraction of these substrates from other sources. The sensitivity<br />

of PCR methods is now such that extraction of DNA and RNA from tissues fixed in<br />

formaldehyde and buffered formalin is considered routine, and we are now able to<br />

extract DNA from ancient tissues, feces, and many other sources. Indeed, in forensic<br />

science, DNA fingerprinting from sources as diverse as residual saliva on food and<br />

microscopic blood deposits is now possible! Indeed, the description of the extraction<br />

of DNA/RNA alone could probably fill several major chapters. It has, however, not<br />

proven desirable or feasible to be exhaustive in our approach to DNA/RNA extraction<br />

protocols, and we have therefore restricted these to major methods in use in many<br />

laboratories. Further references (1–7) that provide detailed reviews of methods for<br />

nucleic acid extraction and some recommended web sites are listed in the reference<br />

section.<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

27


28 <strong>Bartlett</strong><br />

References<br />

1. US Department of Commerce Molecular <strong>Bio</strong>logy Techniques Forum http://research.<br />

nwfsc.noaa.gov/protocols/methods/<br />

2. http://www.stratagene.com/<br />

3. http://www.promega.com/tbs/<br />

4. http://www.highveld.com/protocols.html<br />

5. http://www.dynal.no/<br />

6. Sambrook, J., Fritsch, E. F., and Maniatis, T. (1989) Molecular Cloning: A Laboratory<br />

Manual. Cold Spring Harbor Laboratory Press, New York.<br />

7. <strong>Bartlett</strong>, J. M. S., ed. (2000) Ovarian Cancer: Methods & Protocols, Vol. 39, Methods in<br />

Molecular <strong>Bio</strong>logy. Humana Press, Totowa, NJ.


DNA Extraction from Whole Blood 29<br />

6<br />

Extraction of DNA from Whole Blood<br />

<strong>John</strong> M. S. <strong>Bartlett</strong> and Anne White<br />

1. Introduction<br />

There are many differing protocols and a large number of commercially available<br />

kits used for the extraction of DNA from whole blood. This procedure is one we use<br />

routinely in both research and clinical service provision and is cheap and robust. It<br />

can also be applied to cell pellets from dispersed tissues or cell cultures (omitting<br />

the red blood lysis step.<br />

2. Materials<br />

This method uses standard chemicals that can be obtained from any major supplier;<br />

we use Sigma.<br />

1. Waterbath set at 65°C.<br />

2. Centrifuge tubes (15 mL; Falcon).<br />

3. Microfuge (1.5 mL) tubes.<br />

4. Tube roller/rotator.<br />

5. Glass Pasteur pipets, heated to seal the end and curled to form a “loop” or “hook” for<br />

spooling DNA.<br />

6. EDTA (0.5 M), pH 8.0: Add 146.1 g of anhydrous EDTA to 800 mL of distilled water.<br />

Adjust pH to 8.0 with NaOH pellets (this will require about 20 g). Make up to 1 L with<br />

distilled water. Autoclave at 15 p.s.i. for 15 min.<br />

7. 1 M Tris-HCl, pH 7.6: Dissolve 121.1 g of Tris base in 800 mL of distilled water. Adjust<br />

pH with concentrated HCl (this requires about 60 mL). CAUTION: the addition of acid<br />

produces heat. Allow mixture to cool to room temperature before finally correcting pH.<br />

Make up to 1 L with distilled water. Autoclave at 15 p.s.i. for 15 min.<br />

8. Reagent A: Red blood cell lysis: 0.01M Tris-HCl pH 7.4, 320 mM sucrose, 5 mM MgCl 2 ,<br />

1% Triton X 100.<br />

9. Add 10 mL of 1 M Tris, 109.54 g of sucrose, 0.47 g of MgCl 2 , and 10 mL of Triton X-100<br />

to 800 mL of distilled water. Adjust pH to 8.0, and make up to 1 L with distilled water.<br />

Autoclave at 10 p.s.i. for 10 min (see Note 1).<br />

10. Reagent B: Cell lysis: 0.4 M Tris-HCl, 150 mM NaCl, 0.06 M EDTA, 1% sodium dodecyl<br />

sulphate, pH 8.0. Take 400 mL of 1 M Tris (pH 7.6), 120 mL of 0.5 M EDTA (pH 8.0),<br />

8.76 g of NaCl, and adjust pH to 8.0. Make up to 1 L with distilled water. Autoclave 15 min<br />

at 15. p.s.i. After autoclaving, add 10 g of sodium dodecyl sulphate.<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

29


30 <strong>Bartlett</strong> and White<br />

11. 5 M sodium perchlorate: Dissolve 70 g of sodium perchlorate in 80 mL of distilled water.<br />

Make up to 100 mL.<br />

12. TE Buffer, pH 7.6: Take 10 mL of 1 M Tris-HCl, pH 7.6, 2 mL of 0.5 M EDTA, and make<br />

up to 1 L with distilled water. Adjust pH to 7.6 and autoclave 15 min at 15. p.s.i.<br />

13. Chloroform prechilled to 4°C.<br />

14. Ethanol (100%) prechilled to 4°C.<br />

3. Method<br />

3.1. Blood Collection<br />

1. Collect blood in either a heparin- or EDTA-containing Vacutainer by venipuncture (see<br />

Note 2). Store at room temperature and extract within the same working day.<br />

3.2. DNA Extraction<br />

To extract DNA from cell cultures or disaggregated tissues, omit steps 1 through 3.<br />

1. Place 3 mL of whole blood in a 15-mL falcon tube.<br />

2. Add 12 mL of reagent A.<br />

3. Mix on a rolling or rotating blood mixer for 4 min at room temperature.<br />

4. Centrifuge at 3000g for 5 min at room temperature.<br />

5. Discard supernatant without disturbing cell pellet. Remove remaining moisture by inverting<br />

the tube and blotting onto tissue paper.<br />

6. Add 1 mL of reagent B and vortex briefly to resuspend the cell pellet.<br />

7. Add 250 µL of 5 M sodium perchlorate and mix by inverting tube several times.<br />

8. Place tube in waterbath for 15 to 20 min at 65°C.<br />

9. Allow to cool to room temperature.<br />

10. Add 2 mL of ice-cold chloroform.<br />

11. Mix on a rolling or rotating mixer for 30 to 60 min (see Note 3).<br />

12. Centrifuge at 2400g for 2 min.<br />

13. Transfer upper phase into a clean falcon tube using a sterile pipet.<br />

14. Add 2 to 3 mL of ice-cold ethanol and invert gently to allow DNA to precipitate (see<br />

Note 4).<br />

15. Using a freshly prepared flamed Pasteur pipet spool the DNA onto the hooked end (see<br />

Note 5).<br />

16. Transfer to a 1.5-mL Eppendorf tube and allow to air dry (see Note 6).<br />

17. Resuspend in 200 µL of TE buffer (see Notes 7 and 8).<br />

4. Notes<br />

1. Autoclaving sugars at high temperature can cause caramelization (browning), which<br />

degrades the sugars.<br />

2. As will all body fluids, blood represents a potential biohazard. Care should be taken in all<br />

steps requiring handling of blood. If the subject is from a known high risk category (e.g.,<br />

intravenous drug abusers) additional precautions may be required.<br />

3. Rotation for less than 30 or over 60 min can reduce the DNA yield.<br />

4. DNA should appear as a mucus-like strand in the solution phase.<br />

5. Rotating the hooked end by rolling between thumb and forefinger usually works well.<br />

If the DNA adheres to the hook, break it off into the Eppendorf and resuspend the DNA<br />

before transferring to a fresh tube.<br />

6. Ethanol will interfere with both measurements of DNA concentration and PCR reactions.<br />

However, overdrying the pellet will prolong the resuspension time.


DNA Extraction from Whole Blood 31<br />

7. The small amount of EDTA in TE will not affect PCR. We routinely use 1 µL per PCR<br />

reaction without adverse affects.<br />

8. DNA can be quantified and diluted to a working concentration at this point or simply<br />

use 1 µL per PCR reaction; routinely, we expect 200 to 500 ng/µL DNA to be the yield<br />

of this procedure.


32 <strong>Bartlett</strong> and White


DNA Extraction from Tissue 33<br />

7<br />

DNA Extraction from Tissue<br />

Helen Pearson and David Stirling<br />

1. Introduction<br />

The following protocol is one of the longest-established methods of DNA extraction<br />

and works well with a wide range of solid tissues. Proteins are digested with proteinase<br />

K and extracted with phenol chloroform. DNA is then precipitated with ethanol. The<br />

resultant DNA (10–50 µg) is of high molecular weight and is a suitable template for<br />

long polymerase chain reaction (PCR).<br />

2. Materials<br />

1. Microfuge tubes (1.5 mL).<br />

2. Shaking water bath or incubator with rotisserie.<br />

3. Microfuge.<br />

4. DNA digestion buffer: 50 mM Tris-HCl, 100 mM EDTA, 100 mM NaCl, 1% SDS,<br />

pH 8.0.<br />

5. Proteinase K: 0.5 mg/mL in DNA digestion buffer.<br />

6. Phenol/chloroform/isoamyl alcohol (25241).<br />

7. 100% EtOH.<br />

8. 70% EtOH.<br />

9. TE buffer: 10 mM Tris-HCl, 1 mM EDTA, pH 8.0.<br />

3. Protocol<br />

1. Place 0.1 to 0.5 g of tissue into polypropylene microfuge tube (see Note 1).<br />

2. Add 0.5 mL of DNA digestion buffer with proteinase K (see Note 2).<br />

3. Incubate overnight at 50 to 55ºC with gentle shaking.<br />

4. Spin tubes for 5 s at 500g to collect mix in bottom of tube.<br />

5. Add 0.7 mL of phenol/chloroform/isoamyl alcohol (25241).<br />

6. Mix by inversion for 1 h (do not vortex).<br />

7. Microfuge at 12,000g for 5 min and transfer 0.5 mL of the upper phase to new microfuge<br />

tube.<br />

8. Add 1 mL of 100% ethanol at room temperature and gently invert until DNA precipitate<br />

forms (approx 1 min).<br />

9. Microfuge at 12,000g for 5 min and discard supernatant.<br />

10. Add 1 mL of 70% ethanol (–20ºC) and invert several times. This ethanol wash removes<br />

excess salt, which may otherwise interfere with PCR.<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

33


34 Pearson and Stirling<br />

11. Microfuge at 12,000g for 5 min and discard supernatant.<br />

12. Spin tubes for 5 s to collect any remaining ethanol in bottom of tube. Remove last drops<br />

of ethanol with fine pastette.<br />

13. Air dry at room temperature for 10 to 15 min (any longer will render DNA difficult to<br />

redissolve).<br />

14. Resuspend in 100 µL of TE and incubate at 65°C for 15 min to dissolve DNA (see<br />

Note 3).<br />

4. Notes<br />

1. Some tissues contain large amounts of connective tissue and are difficult to digest. These<br />

can be ground frozen in liquid Nitrogen and ground in a mortar and pestle before being<br />

digested with proteinase K.<br />

2. Proteinase K solution can be kept for several days at 4°C.<br />

3. Repeat pipetting through a narrow gauge tip can help this process.


DNA from Archival Tissues 35<br />

8<br />

Extraction of DNA from Microdissected Archival Tissues<br />

James J. Going<br />

1. Introduction<br />

Many modern analytical methods require little material, and this has made feasible<br />

biochemical and molecular analyses of small tissue fragments, even individual cells,<br />

by microdissection of histological sections (1,2). Polymerase chain reaction (PCR) can<br />

potentially be applied to the analysis of single DNA molecules, as in the analysis of<br />

single haploid cells, such as spermatozoa (3). This sensitivity requires careful attention<br />

to technique and proper controls to avoid false-positive or other spurious results.<br />

Microdissection techniques used by different research groups are diverse, and<br />

recent articles explore different techniques (4–7). This chapter presents a technique of<br />

histological microdissection applicable to a variety of tissues.<br />

Microdissection can be applied to paraffin or frozen sections of human and animal<br />

tissues, depending on availability, but in human studies, it may be necessary to work<br />

with formalin-fixed, paraffin-embedded archival tissues. Although fixed tissues have<br />

disadvantages, particularly the degradation of nucleic acids after fixation, which<br />

may make successful PCR amplification more difficult, better preservation of tissue<br />

morphology compensates. This may be important because one purpose of histological<br />

microdissection is to bring together molecular and morphological analysis of the<br />

same cells. Fixed tissues sections may be easier to handle than unfixed tissues during<br />

microdissection. This chapter concentrates on fixed tissues.<br />

2. Materials<br />

All reagents should be of molecular biology quality.<br />

1. Proteinase K from Tritirachium album (Sigma), 20 mg/mL stock solution. Store 50-µL<br />

aliquots at –20°C, thaw, and dilute to 1 mL with digestion buffer containing 1% Tween to give<br />

working stock solution of proteinase K, 1 mg/mL for tissue digestion to release DNA.<br />

2. Proteinase K digestion buffer, pH 8.3. (TRIS-HCl, 2.2 g/L; TRIS base 4.4 g/L; EDTA<br />

0.37 g/L; separate batches of the buffer should be prepared detergent-free and containing<br />

1% Tween).<br />

3. Leica model M mechanical micromanipulator (other micromanipulators may be suitable).<br />

4. Tungsten wire (0.5 mm in diameter) for dissection needles (or ready-made needles);<br />

bacteriological loop holders for mounting needles in micromanipulator.<br />

5. Facility for electrolytic sharpening of tungsten needles (see Subheading 3.5.)<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

35


36 Going<br />

3. Methods<br />

3.1. Section Cutting<br />

Careful clean techniques should be used when cutting sections for PCR analysis.<br />

1. Use a new part of the microtome blade to cut each section to avoid possible the carryover<br />

of DNA from one tissue block to sections of the next.<br />

2. Ribbons of sections should be floated out on a clean water bath and no buildup of section<br />

debris permitted.<br />

3. Glass slides upon which sections are mounted should be scrupulously clean. Dry mounted<br />

sections at 56°C for 2 h (see Note 1).<br />

3.2. Dewaxing and Staining Sections<br />

Dewax sections completely before microdissection.<br />

1. Immerse 6- to 7-µm sections in a slide rack for 10 min in xylene. Drain surplus xylene<br />

thoroughly from the sections to minimize the carryover of dissolved wax and then transfer<br />

to a second xylene bath for 2 min. Avoid breathing xylene vapor and be aware of the fire<br />

hazards of organic solvents.<br />

2. Take sections through two baths of 99% industrial methylated spirits to remove xylene<br />

and a third bath of 95% industrial methylated spirits.<br />

3. Transfer sections to distilled or deionized water of a satisfactory standard for preparing<br />

PCR reagents. Stain by immersion for 30 s in 0.05% w/v toluidine blue in distilled water<br />

(see Note 2).<br />

4. Wash in distilled water (see Note 3).<br />

5. After dissection, dehydrate and mount slides with a coverslip to provide a permanent<br />

record of the dissection (see Note 4).<br />

3.3. Microdissection Tools<br />

Successful dissection can be conducted freehand using sterile curved scalpel blades<br />

or sterile hypodermic needles, with or without a dissecting microscope.<br />

1. Take a section for dissection from the water bath immediately before it is needed, drain<br />

it, and blot any surplus water from around the section with a disposable, lint-free tissue<br />

(do not touch the section).<br />

2. Stroke the edge of the scalpel blade decisively across the section to remove a strip of tissue,<br />

the width of which will vary with the angle between blade and section (see Note 5).<br />

3. For more precise dissection with the micromanipulator, an electrolytically sharpened<br />

tungsten needle is ideal (Fig. 1). This tool consists of 25 mm of tungsten wire, 0.5 mm<br />

in diameter, sharpened and polished electrolytically to a fine point (tip radius several<br />

microns).<br />

4. Mount the needle in a collet-type bacteriological loop holder.<br />

5. Mount the loop holder in the tool holder of the micromanipulator and angle it downward<br />

25 to 30º.<br />

3.4. Making and Maintaining Tungsten Microneedles<br />

Tungsten needles can be obtained ready-made as ohmic probe needles but are easily<br />

fabricated from plain 0.5-mm tungsten wire (obtainable from suppliers of equipment<br />

for electron microscopy) This requires an electrolytic cell (see Note 6).


DNA from Archival Tissues 37<br />

Fig. 1. Microdissection (stills from a video). Frame 1: the needle is placed at the periphery of<br />

an island of carcinoma cells (colonic carcinoma, paraffin section, toluidine blue stain). Frame<br />

2: about one-quarter of the island of tumor cells has been separated from the glass slide.<br />

Frame 3: the whole island is completely detached, shown by slight rotation with respect to the<br />

position it occupied in the section. The sample is ready for collection, proteinase K digestion,<br />

and further analysis.<br />

1. Obtain ~10 cm of platinum wire (e.g., from an old electrophoresis apparatus) and make<br />

a circular loop in it just small enough to fit inside a standard 20-mL universal container,<br />

with a “tail” 2 to 3 cm long.<br />

2. Drill a 1-mm hole in the side of universal container near the base and thread the platinum<br />

wire tail through it, with the loop neatly placed at the base of the container. Seal the hole<br />

inside and out with an epoxy resin glue, for example, Araldite.<br />

3. Almost fill the cell with 0.1 M potassium hydroxide (KOH; 1.2 g in 20 mL) in water<br />

(caution: KOH is caustic).<br />

4. Connect the platinum wire cathode to the negative (–) terminal of a standard 9-V radio<br />

battery (dry cell), and make the tungsten wire, mounted in its bacteriological loop holder,<br />

the anode (+). Ensure that the cell cannot fall over and spill caustic electrolyte solution.<br />

5. Complete the circuit by dipping 5 to 10 mm of the tungsten wire vertically into the KOH<br />

solution. Hydrogen bubbles will appear at the platinum cathode and nascent oxygen will<br />

remove tungsten from the anode, which sharpens to a fine, polished point. New needles<br />

are made this way and damaged needles refurbished. If sharpening does not occur, check<br />

the polarity.<br />

6. Straighten bent needles by rolling firmly between glass slides, and then polish/resharpen.<br />

7. Rinse needles with distilled water from a wash bottle after sharpening or resharpening to<br />

remove droplets of KOH electrolyte.<br />

8. Store prepared needles in a covered Petri dish with the blunt end pressed lightly into a ring<br />

of modelling clay or similar. Handle with fine forceps.<br />

3.5. Performing Microdissection<br />

1. Retrieve the slide to be dissected from distilled water and dry the back of the slide and<br />

around the section using a clean disposable laboratory tissue.<br />

2. Place the section on the microscope stage (see Note 7) and cover with a pool of proteinase<br />

K lysis buffer (without detergent) from a disposable sterile Pasteur pipet with a rubber-bulb


38 Going<br />

pipet filler. Spread the pool of buffer until it extends 3 to 4 mm beyond all edges of<br />

the section and is as deep as possible without spillage.<br />

3. Center the area of cells in the section to be retrieved for subsequent analysis in the<br />

microscope field at an appropriate magnification.<br />

4. Using the coarse motion controls of the micromanipulator, place the needle over the area<br />

to be dissected.<br />

5. Lower the needle gently at the edge of the area to be dissected until its tip just touches<br />

the slide. Stop lowering the needle when a small lateral deflection of the needle tip occurs<br />

(see Note 8).<br />

6. Microdissection techniques vary for different specimens. In general, attempt by blunt<br />

dissection to develop cleavage between groups of cells. Work around the area to be<br />

dissected, developing a split between the area to be kept and the area to be removed (Fig. 1).<br />

Then, use the point of the needle gradually to undermine the area to be recovered, pushing<br />

and pulling with the tip and side of the needle until the area to be retrieved has been peeled<br />

from the slide and floats freely in the buffer pool.<br />

7. If it is not clear whether the fragment is still attached to the section, gently agitate the<br />

slide. Attached fragments do not move freely.<br />

8. Sometimes a tissue fragment remains attached by a few strands of collagen; a second<br />

tungsten needle in a bacteriological loop holder, used freehand, will often detach it.<br />

3.6. Retrieving Dissected Fragments<br />

1. Bring the tip of the pipet close to the fragment to be retrieved and capture the fragment<br />

by suddenly releasing the pipet plunger, dragging fragment and a fixed volume of buffer<br />

into the pipet tip (see Note 9).<br />

2. Expel the captured specimen into a labeled microcentrifuge tube, ready for further<br />

processing (see Note 8).<br />

3. Check that the microdissected specimen is really in the tube. It helps if the fragment<br />

is easily seen with the naked eye. Use a magnifying lens or the microscope to make<br />

certain. Toluidine blue stained fragments are easy to see; unstained fragments may be<br />

practically invisible.<br />

3.7. Extracting DNA: Proteinase K Digestion<br />

In a fixed specimen, nucleic acids are present in a dense array of crosslinked proteins.<br />

Proteinase K digestion appears effectively to release them and make them available<br />

for subsequent PCR.<br />

1. Add an equal volume of Proteinase K digestion buffer pH 8.3 containing 1 mg/mL of<br />

proteinase K and 1 mg/mL Proteinase K to each specimen tube.<br />

2. Digest microdissected specimen in proteinase K (final concentration 500 µg/mL) at 37°C<br />

overnight in a water bath or incubator (digestion can continue over a weekend without<br />

detriment).<br />

3. Heat specimens in a PCR block (95°C for 10), to inactivate PK.<br />

4. Spin the specimen down by brief centrifugation.<br />

5. Specimens are stable at room temperature for subsequent DNA PCR. Store for longer<br />

periods at 4°C or –20°C.<br />

6. Accurate labeling is crucial (see Note 10).<br />

3.8. DNA Purification after PK Digestion<br />

The 25- or 50-µL sample remaining after PK digestion of a microdissected tissue<br />

fragment contains only small quantities of nucleic acids (DNA, RNA). One thousand


DNA from Archival Tissues 39<br />

cells contain about 6 ng of DNA and will contain at most 2000 copies of an amplifiable<br />

DNA sequence; only 1000 copies of each of two different alleles. An attempt to<br />

purify such quantities of DNA or RNA risks losing the specimen, although techniques<br />

have been described. Such purification does not obviously improve subsequent PCR<br />

amplification. Carefully optimize your PCR using the unpurified digest before you<br />

conclude that such purification is essential. Several published microdissection studies<br />

using the techniques described here have used 1-µL aliquots of sample as template<br />

(8–11). Strategies such as hot start, touchdown, nested, or real-time PCR may to help to<br />

obtain good PCR results by decreasing amplification of spurious products.<br />

Techniques exist for whole-genome amplification of DNA from small samples, even<br />

single cells (12–14), to expand the template pool available for subsequent PCR analysis.<br />

The possibility of introducing artifacts should be considered, but with appropriate<br />

controls may be used for projects with scanty material.<br />

3.9. Frozen vs Paraffin Sections<br />

Frozen sections are less easy to microdissect than paraffin sections. Unfixed are<br />

less robust than fixed tissues and stand up less well to the manipulation necessary<br />

to detach the specimen.<br />

3.10. Analysis of RNA from Microdissected Material<br />

RNA from unfixed tissue may be more suitable for analysis than RNA from fixed<br />

tissue; however, RNA in unfixed tissue may be more susceptible to degradation,<br />

and RT-PCR of RNA from microdissected fixed tissue fragments can be achieved.<br />

Rather than performing the microdissection in a guanidinium-containing buffer, which<br />

removes toluidine blue from the section, performed microdissection in DEPC-treated<br />

distilled water and transfer fragments subsequently to RT-PCR buffers.<br />

4. Notes<br />

1. Drying influences the firmness with which sections adhere to the slide and ease of<br />

dissection. In general, slides coated with poly-L-lysine or treated with silanes should be<br />

avoided because it may be difficult to detach tissue from such slides. Conventional 3- to<br />

4-µm histological sections are a compromise between ease of cutting, depth of staining,<br />

and visibility of cytological detail. For histological microdissection, slightly thicker<br />

sections (6–8 µm) contain more nucleic acid per unit area, and their slightly increased<br />

thickness does not usually cause interpretation problems. With sections over 10 µm, poorer<br />

visualization of cells is a disadvantage. Cut and dried sections should be stored in dust-free<br />

conditions. There seems to be no need to store them in a refrigerator or freezer. Disposable<br />

latex gloves should be worn for all manipulations to reduce contamination.<br />

2. Staining reveals tissue structure, but unstained sections may be dissectible, especially if<br />

a serial hematoxylin and eosin section is available for reference. Staining often makes<br />

dissection easier. Toluidine blue staining is easy and does not seem to interfere with<br />

subsequent PCR, although this should be verified for particular applications.<br />

3. Stained sections can be stored in distilled water until dissection. Refrigeration at 4°C<br />

in water overnight causes some destaining and sections may lift from the slide. Stained<br />

sections can be stored dry but it is best not to dewax and stain more sections than can be<br />

used in a single dissecting session.<br />

4. A serial hematoxylin and eosin section shows what has been removed. A magnified<br />

photocopy or digital image of such a serial section can be annotated on hard copy or


40 Going<br />

digitally to record the dissection. Photocopying histological sections through an acetate<br />

sheet avoids scratching the photocopier glass or smearing it with mounting medium.<br />

5. A scalpel blade vertically in contact with the section will remove a narrow strip of tissue.<br />

The blade at a flatter angle will remove a wider strip. Tissue so removed can often be rolled<br />

into a small ball or pill, picked up on the tip of the scalpel blade or sterile hypodermic<br />

needle and transferred to a microcentrifuge tube for further processing. Tissue is most<br />

easily handled when damp but without excess water, in which dispersed tissue fragments<br />

are hard to retrieve or too dry and brittle. Static electricity may be troublesome.<br />

6. Bacteriological loop holders and tungsten wire are inexpensive. Either prepare enough<br />

needles in advance to use a new needle for each microdissection if necessary or briefly<br />

repolish the needle between samples, which exposes a new tungsten surface. The risk of<br />

DNA carryover on the needle from one specimen to the next appears largely theoretical.<br />

In a study of ras mutation in colorectal carcinomas (10), no false positives were detected<br />

in the analysis of several hundred separate microdissected normal and tumour tissue<br />

samples.<br />

7. Any microscope can be used if there is enough space between objective and stage for<br />

access to the specimen. Ordinary or inverted standard microscopes can be used but a<br />

stereo dissecting microscope with relatively high maximum magnification (up to ×120)<br />

is ideal.<br />

8. If the specimen adheres to the inside of the pipet tip, it can sometimes be dislodged by<br />

repeated in drawing and expulsion of buffer. It may be picked out with a needle, or the<br />

pipet tip may be cut off and placed with the tissue fragment in the tube for subsequent<br />

digestion. Sticking can usually be avoided by coating pipet tips before use with a silicone<br />

such as Sigmacote. Draw (e.g., 50 µL) of Sigmacote into the tip, expel it again, shake<br />

off any excess, and allow to dry. Coat tips an hour or two before you need them. Collect<br />

microdissected fragments for further processing in 500-µL tubes with screw-on lids sealed<br />

with rubber O rings. The O ring prevents loss of specimen volume by evaporation during<br />

subsequent processing and 500-µL tubes fit most PCR thermal cycling blocks for heating<br />

to inactivate Proteinase K after the digestion process is complete<br />

9. Capture volumes of 12.5 or 25 µL work well. The pipet should be nearly vertical to<br />

minimize the risk of damaging the section. Use a Gilson-type pipet with a wide-bore<br />

polypropylene tip. Such tips are in many manufacturer’s catalogs, or you can cut the end<br />

off an ordinary pipet tip. Steady your hand, holding the pipet, against your fist resting on<br />

the microscope stage. The fragment may be hard to retrieve if it lies flat on the section or<br />

slide. Briskly expelling buffer toward the fragment will usually lift the fragment to a level<br />

from which it may easily be recovered. Retrieval of dissected fragments without damage<br />

to the section is easier from a deep pool of buffer. The buffer should not contain detergent,<br />

such as Triton X-100 which, by reducing surface tension, prevents the formation of a<br />

deep standing pool of buffer.<br />

10. Labeling must survive overnight incubation in the water bath and subsequent heat<br />

inactivation in the PCR thermal cycler block, the wells of which often contain oil traces<br />

which may dissolve even permanent marker pen inks. Write specimen numbers on the lid<br />

of the container as well. Typing correction fluid on the lid gives a white surface on which<br />

graphite pencil is permanent. Do not transpose lids. Do not open more than one specimen<br />

tube at a time, and replace the lid at once.<br />

References<br />

1. Schriever F., Freeman G., and Nadler L. M. (1991) Follicular dendritic cells contain a<br />

unique gene repertoire demonstrated by single-cell polymerase chain reaction. Blood<br />

77, 787–791.


DNA from Archival Tissues 41<br />

2. Trumper, L. H., Brady, G., Bagg, A., Gray, D., Loke, S. L., Griesser, H., et al. (1993)<br />

Single-cell analysis of Hodgkin and Reed-Sternberg cells: Molecular heterogeneity of gene<br />

expression and p53 mutations. Blood 81, 3097–3115.<br />

3. Hubert, R., Weber, J. L., Schmitt, K., and Arnheim, N. (1992) A new source of polymorphic<br />

DNA markers for sperm typing: Analysis of microsatellite repeats in single cells. Am. J.<br />

Hum. Genet. 51, 985–991.<br />

4. Emmert-Buck, M. R., Bonner, R. F., Smith, P. D., Chuaqui, R. F., Zhuang, Z. P., Goldstein,<br />

S. R., et al. (1996) Laser capture microdissection. Science 274, 998–1001.<br />

5. Moskaluk, C. A. and Kern, S. E. (1997) Microdissection and polymerase chain reaction<br />

amplification of genomic DNA from histological tissue sections. Am. J. Pathol. 150,<br />

1547–1552.<br />

6. Bohm, M., Wieland, I., Schutze, K., and Rubben, H. (1997) Microbeam MOMeNT—Noncontact<br />

laser microdissection of membrane-mounted native tissue. Am. J. Pathol. 151,<br />

63–67.<br />

7. Fend, F., Emmert-Buck, M. R., Lee, J., Cole, K., Chaqui, R. F., Liotta, L. A., et al. (1998)<br />

Immuno-LCM: laser capture microdissection of immunostained frozen sections for RNA<br />

analysis. Mod. Pathol. 11, A185.<br />

8. Murphy, D. S., Hoare, S. F., Going, J. J., Mallon, E. A., George, W. D., Kaye, S. B., et<br />

al. (1995) Characterization of extensive genetic alterations in ductal carcinoma in situ<br />

by fluorescence in situ hybridization and molecular analysis. J. Natl. Cancer Inst. 87,<br />

1694–1704.<br />

9. Going, J. J. and Lamb, R. F. (1996) Practical histological dissection for PCR analysis.<br />

J. Pathol. 179, 121–124.<br />

10. Lamb, R. F., Going, J. J., Pickford, I., and Birnie, G. D. (1996) Allelic imbalance at the<br />

NME1 locus in microdissected primary and metastatic human colorectal carcinomas is<br />

frequent, but not associated with metastasis to lymph nodes or liver. Cancer Res. 56,<br />

916–920.<br />

11. Al Mulla, F., Going, J. J., Sowden, E. T. H. H., Winter, A., Pickford, I. R., and Birnie,<br />

G. D. (1998) Heterogeneity of mutant versus wild-type Ki-ras in primary and metastatic<br />

colorectal carcinomas, and association of codon-12 valine with early mortality. J. Pathol.<br />

185, 130–138.<br />

12. Zhang, L., Cui, X., Schmitt, K., Hubert, R., Navidi, W., and Arnheim, N. (1992) Whole<br />

genome amplification from a single cell: implications for genetic analysis. Proc. Natl.<br />

Acad. Sci. USA 89, 5847–5851.<br />

13. Telenius, H., Carter, N. P., Bebb, C. E., Nordenskjold, M., Ponder, B. A. J., and Tunnacliffe,<br />

A. (1992) Degenerate oligonucleotide-primed PCR: General amplification of target DNA<br />

by a single degenerate primer. Genomics 13, 718–725.<br />

14. Grothues, D., Cantor, C. R., and Smith, C. L. (1993) PCR amplification of megabase DNA<br />

with tagged random primers (T-PCR). Nucleic Acids Res. 5, 1321–1322.


42 Going


RNA Extraction from Blood 43<br />

9<br />

RNA Extraction from Blood<br />

Helen Pearson<br />

1. Introduction<br />

Based on the method of Chomczynski and Sacchi (1), this is an extremely reliable<br />

method without the requirement for centrifugation over CsCl gradients. As with any<br />

RNA protocol, extreme care should be taken to exclude RNAse contamination, the<br />

greatest source of which will be the sample itself. All disposables and reagents should<br />

be RNAse free.<br />

2. Materials<br />

1. Microfuge tubes (1.5 mL).<br />

2. Ice bucket.<br />

3. Microfuge.<br />

4. Red cell lysis buffer: 1.6 M sucrose, 5% Triton X-100, 25 mM MgCl 2 , 60 mM Tris-HCl,<br />

pH 7.5; stored at 2–8°C and used cold.<br />

5. Extraction Buffer: 5.25 M guanidinium thiocyanate, 50 mM Tris-Cl, pH. 6.4, 20 mM<br />

EDTA, 1% Triton X-100, 0.1 M β-mercaptoethanol (add immediately prior to use).<br />

6. 2 M sodium acetate, pH 4.0.<br />

7. Phenol (saturated with 1 M Tris-HCl: 0.1 M EDTA, pH 8.0).<br />

8. Chloroform:Iso-amyl alcohol (241).<br />

9. Isopropyl alcohol.<br />

10. 70% Ethanol.<br />

11. RNAse-free distilled water.<br />

3. Method<br />

1. In a microfuge tube, mix 100 µL anticoagulated blood with 1 mL of red cell lysis buffer<br />

(see Notes 1–3).<br />

2. Leave at room temperature with occasional shaking until the red cells have lysed and the<br />

solution translucent (usually within 5 min).<br />

3. Microfuge for 30 s at 13,000g to pellet the white blood cells. Remove and discard<br />

supernatant.<br />

4. Add 200 µL of extraction buffer and resuspend cell pellet by drawing through narrow<br />

gauge needle several times.<br />

5. Add 20 µL of 2 M sodium acetate and mix gently by inversion.<br />

6. Add 220 µL of phenol and mix gently by inversion.<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

43


44 Pearson<br />

7. Add 60 µL of chloroform/isoamyl alcohol (241) and vortex vigorously.<br />

8. Place on ice for 15 min.<br />

9. Microfuge at 12,000g for 5 min and transfer the upper phase to new microfuge tube.<br />

10. Add 200 µL of ice-cold isopropanol mix and store at –20°C for 30 min.<br />

11. Microfuge at 12,000g for 15 min and discard supernatant.<br />

12. Resuspend pellet in 200 µL of extraction buffer.<br />

13. Repeat steps 3 through 9.<br />

14. Wash pellet with 400 µL of cold 70% ethanol.<br />

15. Microfuge at 12,000g for 5 min and discard supernatant.<br />

16. Carefully remove last traces of ethanol from tube (folded sterile swab or kimwipe works<br />

well).<br />

17. Resuspend in 100 µL of distilled water and incubate at 50°C for 15 min to dissolve RNA<br />

(see Note 4).<br />

4. Notes<br />

1. Blood stored at room temperature or 4°C should be mixed thoroughly prior to aliquots<br />

being removed.<br />

2. Frozen blood samples should be allowed to thaw completely and mixed thoroughly before<br />

aliquots being removed. Although freezing lyses red blood cells, the red cell lysis step<br />

should still be performed to efficiently remove hemoglobin from the sample. Repeated<br />

freeze/thaw cycles should be avoided.<br />

3. Buffy coat contains two to four times the amount of white blood cells per volume compared<br />

to fresh blood. Therefore, it is advisable to use only 50 µL of buffy coat diluted with 50 µL<br />

of phosphate-buffered saline as starting material for this protocol.<br />

4. Repeat pipetting through a narrow gauge tip can help this process.<br />

References<br />

1. Chomczynski, P. and Sacchi. N. (1987) Single-step method of RNA isolation by acid<br />

guanidinium thiocyanate-phenol-chloroform extraction. Anal. <strong>Bio</strong>chem. 162, 156–159.


RNA Extraction from Frozen Tissue 45<br />

10<br />

RNA Extraction from Frozen Tissue<br />

<strong>John</strong> M. S. <strong>Bartlett</strong><br />

1. Introduction<br />

RNA extraction is fundamental to all aspects of mRNA analysis. We include here a<br />

simple method that avoids the use of a mortar and pestle.<br />

2. Materials<br />

All chemicals, unless otherwise noted, were molecular biology grade and obtained<br />

from Sigma UK (Poole, Dorset). All glassware was pretreated with di-ethylpyrocarbonate<br />

(DEPC). All deionized distilled water was pretreated with DEPC and autoclaved<br />

(DEPC water). DEPC is a potent anti-RNAse agent.<br />

2.1. DEPC Treatment of Glassware/Distilled Water<br />

0.1% DEPC was added to distilled deionized water and glassware filled and left to<br />

stand overnight. The water was decanted and autoclaved (DEPC-treated water) and<br />

glassware sterilized at 220°C for 2 h (DEPC-treated glassware). DEPC is driven off<br />

by both procedures.<br />

2.2. RNA Extraction<br />

1. Braun Microdismembranator and Teflon vessels (Braun GmBH, Germany).<br />

2. 3 M lithium chloride/6 M urea: Dissolve in 800 mL of DEPC water and make up to 1 L.<br />

The solution can be stored at 4°C for 3 to 6 mo.<br />

3. 10 mM Tris-HCl, 0.5% sodium dodecyl sulphate (SDS) pH 7.5. Prepare stock solutions<br />

of 10% SDS and 0.5 M Tris-HCl (pH 7.5) in DEPC water. Stocks are stable at room<br />

temperature for up to 12 mo.<br />

4. Proteinase K: Prepare 1 mg/mL w/v DEPC water stock and store at –20°C for up to 12 mo.<br />

Dilute in 10 mM Tris-HCl, 0.5% SDS as required, discard unused diluted enzyme.<br />

5. Phenolchloroformisoamyl alcohol: phenol is presaturated with 10 mM Tris-HCl,<br />

pH 7.5. Prepare a mixture of 25241 phenolchloroformisoamyl alcohol (v/v/v). Store<br />

at room temperature for up to 6 mo, shielding from light.<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

45


46 <strong>Bartlett</strong><br />

3. Methods<br />

3.1. RNA Extraction<br />

1. Tissues should ideally be collected fresh and stored in liquid nitrogen. Routinely samples<br />

are collected on ice and transported for freezing within 30 to 60 min.<br />

2. Tissues are disaggregated using a Braun-micro dismembranator. Teflon vessels and steel<br />

ball bearings are cooled in liquid nitrogen before use. Frozen tissue (50–500 mg) is placed<br />

in the vessel with a single ball bearing and agitated at 1000 cycles/second for 60 s. The<br />

vessel is then re-cooled in liquid nitrogen. This process is repeated until tissue is powdered<br />

(usually 2×; see Note 1).<br />

3. Immediately after disaggregation of tissue, tissue material is resuspended while frozen<br />

by adding 1.5 mL of LiCl/Urea and transfer to a separate tube. The vessel is washed a<br />

further 2× with 1.5 mL of LiCl/Urea and the washing combined with the original sample.<br />

The resuspended medium is made up to 6 mL in LiCl/Urea and sonicated for 2× 30 s<br />

at maximum power using a probe sonicator. The sonicated samples are stored overnight<br />

at 4°C (see Note 2).<br />

4. Centrifuge at 15,000g, 4°C for 30 min. The supernatant is discarded and the pellet washed<br />

with a further 6 mL of lithium chloride/urea, recentrifuged (15,000g, 4°C for 30 min) and<br />

the supernatant again discarded.<br />

5. The pellet is resuspended in 6 mL of Tris-HCl/SDS with 50 µg/mL proteinase K (Boehringer<br />

Mannheim, UK) and incubated at 37°C for 20 min.<br />

6. Samples are extracted with 100% phenol, followed by phenol:chloroform:isoamyl-alcohol.<br />

After each extraction, the sample is centrifuged at 2000g at room temperature for 10 min<br />

and the aqueous phase recovered.<br />

7. After the final extraction, 300 µL of 8 M LiCl and 2.5 volumes absolute alcohol are added<br />

and samples stored at –20°C for 30 min overnight. RNA is pelleted by centrifugation<br />

at 4000g, 4°C for 45 min. The supernatant is discarded and the RNA pelleted dried and<br />

resuspended in DEPC-treated distilled water. Concentrations are estimated by optical<br />

density at 260/280 nm.<br />

4. Notes<br />

1. Disaggregation is critically dependent on tissue structure. Most tissues are readily<br />

disaggregated in two 60-s bursts. Other tissue types (e.g., fibrous tissues) may require<br />

longer periods to disrupt tissue. If a mechanical dismembranator is not available, other<br />

methods of tissue homogenization work equally well, either using a mortar and pestle<br />

or blade homogenizers.<br />

2. Other methods can be used to lyse cells, such as passage through a syringe needle, etc.<br />

Extraction of RNA from solid tissues can be problematic because many of the commercial<br />

systems available for RNA extraction are validated for extraction of RNA from cell<br />

culture material or blood lymphocytes. These kits have often been less successful with<br />

tissue-derived material.


RNA Extraction from Tissue Sections 47<br />

11<br />

RNA Extraction from Tissue Sections<br />

Helen Pearson<br />

1. Introduction<br />

There are two different methods of preparing tissue for histology: paraffin-embedding<br />

and freeze-embedding. Each has their advantages and drawbacks. Paraffin-embedded<br />

tissues (PET) produce optimum morphology but have comparatively poor molecular<br />

preservation and recovery. Although frozen sections have poorer histology, they allow<br />

excellent recovery of DNA and RNA for analysis.<br />

Although fixation is performed to preserve the morphology of the living tissue, it<br />

does not necessarily have a beneficial effect on the DNA and RNA. Formalin, one of the<br />

most popular fixatives, crosslinks nucleic acids to protein, thus making the molecules<br />

rigid and susceptible to mechanical shearing. The duration of formalin fixation also<br />

seems to be important. Studies that have demonstrated DNA recovery around 200 base<br />

pairs recommend a period of fixation from 16 to 24 h but not any longer (1).<br />

RNA is a more labile species, and the paraffin-embedding process has been shown<br />

to greatly harm it. Many studies have shown that formalin fixation has the worst effects<br />

among commonly used fixatives and ethanol-based fixatives as having the best RNA<br />

preservation.<br />

2. Materials<br />

1. Microfuge tubes (1.5 mL).<br />

2. Ice bucket.<br />

3. Microfuge.<br />

4. Extraction Buffer: 5.25 M guanidinium thiocyanate, 50 mM Tris-HCl, pH. 6.4, 20 mM<br />

EDTA, 1% Triton X-100, 0.1 M β-mercaptoethanol (add immediately before use.<br />

5. Glycogen (10 mg/mL) in distilled water.<br />

6. 2 M sodium acetate, pH 4.0.<br />

7. Phenol (saturated with 1 M Tris-HCl, 0.1 M EDTA, pH 8.0).<br />

8. Chloroformiso-amyl alcohol (241).<br />

9. Isopropyl alcohol.<br />

10. 70% ethanol.<br />

11. RNAse-free distilled water.<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

47


48 Pearson<br />

3. Method<br />

The method of Chomczynski and Sacchi (2) described in the previous protocol<br />

works well for fixed tissues with only two modifications.<br />

1. The extraction is initiated by incubating tissue sections or microdissected cells in 500-µL<br />

extraction buffer for 5 min at room temperature with gentle agitation inverting several<br />

times.<br />

2. Add 20 µL of 2 M sodium acetate and mix gently by inversion.<br />

3. Add 220 µL of phenol and mix gently by inversion.<br />

4. Add 60 µL of chloroform/isoamyl alcohol (241) and vortex vigorously.<br />

5. Place on ice for 15 min.<br />

6. Microfuge at 12,000g for 5 min and transfer the upper phase to new microfuge tube.<br />

7. Add 1 to 2 µL of glycogen (10 mg/mL). Glycogen is a carrier that is used if RNA quantities<br />

are less than 1 µg. It also facilitates visualization of the pellet.<br />

8. Add 200 µL of ice-cold isopropanol mix and store at –20°C for 30 min.<br />

9. Microfuge at 12,000g for 15 min and discard supernatant.<br />

10. Resuspend pellet in 200 µL of extraction buffer.<br />

11. Repeat steps 3 through 9.<br />

12. Wash pellet with 400 µL of cold 70% ethanol.<br />

13. Microfuge at 12,000g for 5 min and discard supernatant.<br />

14. Carefully remove last traces of ethanol from tube (folded sterile swab or kimwipe works<br />

well).<br />

15. Resuspend in 100 µL of RNAse free distilled water and incubate at 50°C for 15 min to<br />

dissolve RNA (see Note 1).<br />

4. Notes<br />

1. Repeat pipetting through a narrow gauge tip can help this process.<br />

References<br />

1. Foss, R. D., Guha-Thakurta, N., Conran, R. M., and Gutman, P. (1994) Effects of fixative<br />

and fixation time on the extraction and polymerase chain reaction amplification of RNA<br />

from paraffin-embedded tissue. Diagn. Mol. Pathol. 3, 148–155.<br />

2. Chomczynski, P. and Sacchi, N. (1987) Single-step method of RNA isolation by acid<br />

guanidinium thiocyanate-phenol-chloroform extraction. Anal. <strong>Bio</strong>chem. 162, 156–159.


Dual DNA/RNA Extraction 49<br />

12<br />

Dual DNA/RNA Extraction<br />

David Stirling and <strong>John</strong> M. S. <strong>Bartlett</strong><br />

1. Introduction<br />

It is sometimes desirable to extract both RNA and DNA from the same sample,<br />

especially when the sample is small. This can be achieved by isolating a total nucleic<br />

acid fraction that is then divided into two portions, which are treated differentially with<br />

either Dnase I (to remove DNA and recover RNA) or with RNase A (to selectively<br />

recover the DNA); however, this wastes half of the DNA and RNA. An alternative<br />

approach is to sequentially isolate the RNA and DNA fractions from the same sample.<br />

This protocol based on one reported by Chevillard (1), begins by extracting RNA as in<br />

Chapter 9, but then re-extracts the DNA from the collected organic phases. The method<br />

described is for the extraction of both DNA/RNA from tissue but can be modified for<br />

either blood or cell lines (see Notes 1 and 2).<br />

2. Materials<br />

All chemicals, unless otherwise noted, are molecular biology grade and obtained<br />

from Sigma U.K. (Poole, Dorset). All glassware was pretreated with di-ethylpyrocarbonate<br />

(DEPC). All deionized distilled water was pretreated with DEPC and autoclaved<br />

(DEPC water). DEPC is a potent anti-RNAse agent.<br />

2.1. DEPC Treatment of Glassware/Distilled Water<br />

0.1% DEPC is added to distilled deionized water and glassware filled and left to<br />

stand overnight. The water was decanted and autoclaved (DEPC-treated water) and<br />

glassware sterilized at 220°C for 2 h (DEPC-treated glassware). DEPC is driven off<br />

by both procedures.<br />

2.2. RNA Extraction<br />

1. Braun Microdismembranator and Teflon vessels (Braun GmBH, Germany).<br />

2. 3 M lithium chloride/6 M urea: Dissolve in 800 mL of DEPC water and make up to 1 L.<br />

This solution can be stored at 4°C for 3 to 6 mo.<br />

3. 10 mM Tris-HCl, 0.5% sodium dodecyl sulphate (SDS), pH 7.5. Prepare stock solutions of<br />

10% SDS, 0.5 M Tris-HCl (pH 7.5) in DEPC water. Stocks are stable at room temperature<br />

for up to 12 mo.<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

49


50 Stirling and <strong>Bartlett</strong><br />

4. Proteinase K: Prepare 1 mg/mL w/v DEPC water stock, which can be stored at –20°C<br />

for up to 12 mo. Dilute in 10 mM Tris-Cl/0.5% SDS as required, and discard unused<br />

diluted enzyme.<br />

5. Phenol/chloroform/isoamyl alcohol: Phenol is presaturated with 10 mM Tris-HCl, pH 7.5.<br />

Prepare a mixture of 25241 phenolchloroformisoamyl alcohol (v/v/v). Store at room<br />

temperature for up to 6 mo, shielded from light.<br />

6. TE Buffer, pH 7.6: take 10 mL of 1 M Tris-HCl, pH 7.6, 2 mL of 0.5 M EDTA, and make<br />

up to 1 L with distilled water. Adjust pH to 7.6. Autoclave 15 min at 15. p.s.i.<br />

2.3. DNA Extraction (as for RNA Plus)<br />

1. Dual extraction buffer: 0.1 M NaCl, 10 mM Tris-HCl, pH 8.0, 1 mM EDTA, 1% SDS.<br />

Adjust to pH 12.0 with 5 N NaOH immediately before use.<br />

2. 7.5 M ammonium acetate.<br />

3. Methods<br />

3.1. RNA Extraction<br />

1. Tissues should ideally be collected fresh and stored in liquid nitrogen. Routinely samples<br />

are collected on ice and transported for freezing within 30 to 60 min.<br />

2. Tissues are disaggregated using a Braun-micro dismembranator. Teflon vessels and steel<br />

ball bearings are cooled in liquid nitrogen before use. Frozen tissue (50 to 500 mg) is<br />

placed in the vessel with a single ball bearing and agitated at 1000 cycles/second for 60 s.<br />

The vessel is then recooled in liquid nitrogen. This process is repeated until tissue is<br />

powdered (usually twice; see Note 3).<br />

3. Immediately after disaggregation of tissue, material is resuspended while frozen in 1.5 mL of<br />

LiCl/Urea and transferred to a separate tube. The vessel is washed a further 2× with 1.5 mL<br />

of LiCl/Urea and the washing combined with the original sample. The resuspended medium<br />

is made up to 6 mL in LiCl/Urea and sonicated for 2× 30 s at maximum power using a<br />

probe sonicator. The sonicated samples are stored overnight at 4°C (see Note 4).<br />

4. Centrifuge at 15,000g, 4°C for 30 min. The supernatant is discarded and the pellet washed<br />

with a further 6 mL of lithium chloride/urea, recentrifuged (15,000g at 4°C for 30 min)<br />

and the supernatant again discarded.<br />

5. The pellet is resuspended in 6 mL of Tris-HCl/SDS with 50 µg/mL proteinase K (Boehringer<br />

Mannheim, UK), and incubated at 37°C for 20 min.<br />

6. Samples are mixed with an equal volume of phenolchloroformisoamyl-alcohol and<br />

mixed by inversion several times.<br />

7. After mixing, the sample is centrifuged at 2000g at room temperature for 10 min and<br />

the aqueous phase recovered for RNA extraction, the organic phase is retained for DNA<br />

extraction.<br />

8. Repeat steps 6 and 7.<br />

9. After the final extraction, 300 µL of 8 M LiCl and 2.5 volumes absolute alcohol are added<br />

and samples stored at –20°C for 30 min overnight. RNA is pelleted by centrifugation at<br />

4000g, 4°C for 45 min.<br />

10. The supernatant is discarded and the RNA pelleted dried and resuspended in 50 µL of TE.<br />

Concentrations are estimated by optical density at 260/280 nm.<br />

3.2. DNA Extraction<br />

1. Combine the organic phases, including the interfaces from step 7 above (both times), in<br />

a 15-mL polypropylene tube.<br />

2. Add an equal volume of extraction buffer, vortex for 1 min, and place on ice for 10 min.


Dual DNA/RNA Extraction 51<br />

3. Centrifuge for 20 min at 10,000g 4°C.<br />

4. Transfer aqueous phase to fresh tube and add 1/15 volume 7.5 M NH 4 OAc and 2 volumes<br />

ice-cold EtOH. Incubate at –20°C for at least 1 h.<br />

5. Centrifuge for 20 min at 10,000g, 4°C.<br />

6. Carefully decant the supernatant and wash the pellet with 1 mL of 70% ethanol.<br />

7. Centrifuge briefly to ensure that the pellet remains attached.<br />

8. Carefully remove the supernatant and air dry pellet for 10 to 15 min.<br />

9. Resuspend the DNA pellet in 50 µL of TE. Heat 5 min at 55°C then vortex thoroughly<br />

to dissolve the DNA.<br />

4. Notes<br />

1. For extraction from cell lines, scrape cells into 1.5 mL of lithium chloride/urea and then<br />

proceed from step 3 of the RNA extraction above.<br />

2. Extraction of DNA/RNA from blood can be achieved by collecting blood in either a heparin<br />

or EDTA containing Vacutainer by venipuncture. Store at room temperature and extract<br />

within the same working day. Whole blood (3 mL) is placed in a 15-mL polypropylene<br />

tube and mixed with 12 mL of red blood cell lysis solution (0.01 M Tris, pH 7.4, 320 mM<br />

sucrose, 5 mM MgCl 2 , 1% Triton X 100). Blood is then mixed on a rolling or rotating blood<br />

mixer for 4 min at room temperature. Lymphocytes are recovered by centrifugation at<br />

3000g for 5 min at room temperature. Then, proceed with RNA extract at step 3 above.<br />

3. Disaggregation is critically dependent on tissue structure. Most tissues are readily<br />

disaggregated in two 60-s bursts. Other tissue types (e.g., fibrous tissues) may require<br />

longer periods to disrupt tissue. If a mechanical dismembranator is not available, other<br />

methods of tissue homogenization work equally well, either using a mortar and pestle<br />

or blade homogenizers.<br />

4. Other methods can be used to lyse cells, such as passage through a syringe needle, etc.<br />

Extraction of RNA from solid tissues can be problematic because many of the commercial<br />

systems available for RNA extraction are validated for extraction of RNA from cell<br />

culture material or blood lymphocytes. These kits have often been less successful with<br />

tissue-derived material.<br />

References<br />

1. Chevillard, S. (1993) A method for sequential extraction of RNA and DNA from the same<br />

sample, specially designed for a limited supply of biological material. <strong>Bio</strong>Techniques<br />

15, 22–24.


52 Stirling and <strong>Bartlett</strong>


DNA Extraction from Fungi, Yeast, Bacteria 53<br />

13<br />

DNA Extraction from Fungi, Yeast, and Bacteria<br />

David Stirling<br />

1. Introduction<br />

Although individual microorganisms may well require a unique DNA extraction<br />

procedure, here we include robust techniques for the preparation of DNA from fungi,<br />

yeast, and bacteria, which yield DNA suitable for a PCR template.<br />

2. Materials<br />

2.1. Fungal Extraction<br />

1. CTAB extraction buffer: 0.1 M Tris-HCl, pH 7.5, 1% CTAB (mixed hexadecyl trimethyl<br />

ammonium bromide), 0.7 M NaCl, 10 mM EDTA, 1% 2-mercaptoethanol. Add proteinase<br />

K to a final concentration of 0.3 mg/mL prior to use.<br />

2. Chloroformisoamyl alcohol (241).<br />

2.2. Yeast Extraction<br />

1. Yeast extraction buffer A: 2% Triton X-100, 1% sodium dodecyl sulphate, 100 mM<br />

NaCl, 10 mM Tris-HCl, pH 8.0, 1 mM EDTA, pH 8.0. Phenolchloroformisoamyl<br />

alcohol: Phenol is presaturated with 10 mM Tris-HCl, pH 7.5. Prepare a mixture of<br />

25241 phenolchloroformisoamyl alcohol (v/v/v). This solution can be stored at room<br />

temperature for up to 6 mo, shielded from light.<br />

2. Glass beads. Diameter range 0.04–0.07 mm (Jencons Scientific Ltd, UK), suspended as<br />

500 mg/mL slurry in distiller water.<br />

3. Ammonium acetate (4 M).<br />

2.3. Bacterial DNA Protocol<br />

1. Lysozyme/RNase mixture: 10 mg/mL lysozyme, 1 mg/mL RNase, 50 mM Tris-HCl<br />

(pH 8.0). Store at –20°C in small aliquots. Do not refreeze after thawing.<br />

2. STET: 8% sucrose, 5% Triton X-100, 50 mM Tris-HCl (pH 8.0), 50 mM EDTA, pH 8.0.<br />

3. Filter sterilize and store at 4°C.<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

53


54 Stirling<br />

3. Methods<br />

3.1. Fungal Protocol<br />

1. Grind 0.2 to 0.5 g (dry weight) of lyophilized mycellar pad in a mortar and pestle. Transfer<br />

to a 50-mL disposable centrifuge tube.<br />

2. Add 10 mL (for a 0.5 g pad) of CTAB extraction buffer.<br />

3. Gently mix to wet all the powdered pad.<br />

4. Place in 65°C water bath for 30 min.<br />

5. Cool and add an equal volume of chloroform/isoamyl alcohol (241).<br />

6. Mix and centrifuge at 2000g for 10 min at room temperature.<br />

7. Transfer aqueous supernatant to a new tube.<br />

8. Add an equal volume of isopropanol.<br />

9. High molecular weight DNA should precipitate upon mixing and can be spooled out with<br />

a glass rod or hook.<br />

10. Rinse the spooled DNA with 70% ethanol.<br />

11. Air dry, add 1 to 5 mL of TE containing 20 µg/ mL RNAse A. To resuspend the samples,<br />

place in 65°C bath or allow pellets to resuspend overnight at 4°C.<br />

3.2. Yeast Protocol<br />

1. Collect cells from fresh 5 mL culture by centrifugation at 2000g for 10 min and resuspend<br />

in 0.5 mL of water.<br />

2. Transfer cells to 1.5-mL microfuge tube and collect by centrifugation at 15,000g for 10 min,<br />

pour off supernatant and resuspend in residual liquid.<br />

3. Add 0.2 mL of buffer A, 200 µL of glass beads, and 0.2 mL of phenolchloroformisoamyl<br />

alcohol (25241).<br />

4. Vortex for 3 min and add 0.2 mL of TE.<br />

5. Centrifuge at 15,000g for 5 min and then transfer aqueous to new tube.<br />

6. Add 1 mL of 100% EtOH (room temperature), invert tube to mix, and centrifuge at<br />

15,000g for 2 min.<br />

7. Discard supernatant and resuspend pellet in 0.4 mL of TE (no need to dry pellet).<br />

8. Add 10 µL of 4 M ammonium acetate, mix, and then add 1 mL of 100% EtOH and mix.<br />

9. Centrifuge at 15,000g for 2 min and dry pellet. Resuspend in 50 µL of TE.<br />

3.3. Bacterial DNA Protocol<br />

1. Collect the bacteria from a 15-mL overnight culture into a 1.5-mL microfuge tube.<br />

2. Resuspend pellet with 300 µL of STET buffer and add 30 µL of RNAse/lysozyme<br />

mixture.<br />

3. Boil for 1 min 15 s.<br />

4. Centrifuge at 15,000g for at least 15 min.<br />

5. Take supernatant and phenol extract with 150 µL of STET-saturated phenol.<br />

6. Spin and take supernatant. Add 1/10 volume 4 M lithium chloride (autoclaved). Let sit<br />

on ice for 5 to 10 min.<br />

7. Spin and take supernatant. Add equal volume isopropanol at room temperature and<br />

incubate for 5 min.<br />

8. Centrifuge at 15,000g for at least 15 min. No pellet will be visible.<br />

9. IMPORTANT: wash with 80% ethanol (95% will cause the residual Triton to precipitate).<br />

10. Resuspend pellet in 50 to 200 µL of TE.


Isolation of RNA Viruses 55<br />

14<br />

Isolation of RNA Viruses from <strong>Bio</strong>logical Materials<br />

Susan McDonagh<br />

1. Introduction<br />

The successful extraction of viral RNA from biological material requires rapid<br />

transport and adequate storage of samples because of the unstable nature of RNA.<br />

Samples should be received and processed within 6 h and the relevant fractions stored at<br />

–70°C until testing. Also, it is difficult to ascertain the efficiency of sample preparation<br />

methods; therefore, known standards should be processed alongside samples to assess<br />

the loss within the system (1), particularly for quantitative applications.<br />

2. Materials<br />

Unless stated, all chemicals are supplied by Sigma, Poole, UK, or Merck. All stock<br />

solutions should be made using RNA-free water.<br />

2.1. Sample Preparation<br />

1. TNE-buffer: 0.11 M NaCl, 55 mM Tris (pH 8.0), 1.1 mM EDTA pH 8.0, 0.55% sodium<br />

dodecyl sulphate [SDS].<br />

2. Poly-adenylic acid 2 mg/mL (Poly-A; Pharmacia, Upsala, Sweden).<br />

3. Proteinase K, 10 mg/mL in water.<br />

4. Phenol.<br />

5. Phenolchloroform (11).<br />

6. Chloroformisoamylalcohol, 501 (v/v).<br />

7. 3 M sodium acetate, pH 5.2.<br />

8. Ethanol 100% and 80%.<br />

9. 10% SDS.<br />

3. Methods<br />

Because an ultrasensitive system is required, it is necessary to include an ultracentrifugation<br />

step where serum or plasma is concentrated by centrifugation at 27,000g<br />

for 1 h. Large volumes of 1 mL or greater can be used, and most of the supernatant<br />

(900 µL) can be removed before resuspending the pellet in the remaining 100 µL<br />

(see Notes 1 and 2).<br />

1. Prepare a solution of TNE buffer, 0.5% SDS, 1 mg/mL proteinase K, and 40 µg/mL poly<br />

A. Pre-incubate the solution for 10 min at 37°C to inactivate RNases.<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

55


56 McDonagh<br />

2. Add 0.4 mL of TNE buffer with SDS, proteinase K, and poly A to the extraction tubes.<br />

3. Add 100 µL of concentrated plasma or serum and mix immediately.<br />

4. Incubate the lysates for 1.5 to 2 h in 37°C water bath.<br />

5. Phenol extraction: Add 450 µL of phenol to the extraction tubes. Mix extensively and<br />

centrifuge at 13,000g for 10 min.<br />

6. Phenol/chloroform extraction: Transfer the upper aqueous layer to a fresh tube and add<br />

0.45 mL of phenolchloroform (11). Vortex and centrifuge as above.<br />

7. Chloroform/isoamylalcohol extraction: Transfer the upper aqueous layer to a fresh tube<br />

and add chloroformisoamylalcohol (501). Vortex and centrifuge as above.<br />

8. Ethanol precipitation: Transfer the aqueous layer to a fresh tube containing 40 µL of 3 M<br />

Na-acetate, pH 5.2, and 800 µL of ethanol. Mix and precipitate the nucleic acids at –20°C<br />

overnight or at –70°C for 30 min.<br />

9. Collect the nucleic acid by centrifugation at 15,000g for 20 min at 0°C.<br />

10. Discard the supernatant and wash the pellet with 1 mL 80% (v/v) ethanol.<br />

11. Dry the precipitate on a dry block and dissolve in 25 µL of RNase-free water. These<br />

extracts can be stored at –70°C.<br />

4. Notes<br />

1. It is important to process the samples alongside any controls and standards to allow<br />

for the loss of RNA through sample preparation methods. This is especially important<br />

when preparing template for a limiting dilution curve or when using an external curve<br />

dilution series.<br />

2. Although several commercial extraction systems have become available, we have not<br />

attained similar levels of sensitivity as with the protocol suggested here. However,<br />

extraction systems using guanidinium thiocyanate alongside phenol/chloroform or silica<br />

have been successfully used in ultrasensitive assays.<br />

References<br />

1. Clementi, M., Menzo, S., Bagnarelli, P., Manzin, A., Valenza, A., and Varaldo, P. E. (1993)<br />

Quantitative PCR and RT-PCR in virology. PCR Meth. Appl. 2, 191–196.


Extraction of Ancient DNA 57<br />

15<br />

Extraction of Ancient DNA<br />

Wera M. Schmerer<br />

1. Introduction<br />

The DNA extraction process represents one of the critical stages in the analysis<br />

of degraded or ancient DNA. If polymerase chain reaction (PCR) amplification starts<br />

from a poor extract containing low template quantities, stochastic variation in the<br />

amplification of individual alleles may lead to allelic dropout (1), resulting in a high<br />

risk of false-homozygous typing of a heterozygous sample (2).<br />

In analyzing repetitive sequences, such as STR loci, besides quantity, the quality<br />

of the extracted DNA used as template is of particular importance. PCR amplification<br />

of STR loci is generally accompanied by the generation of so-called shadow bands,<br />

byproducts that are shortened in length by one repeat unit compared with the allelic<br />

product (3). Because the accumulation of this artifact is increased with degradation of<br />

the DNA template (4) when amplifying ancient DNA, the intensity of a shadow band<br />

can exceed the intensity of the allelic product. As “artifact alleles” (2,5), these products<br />

may be mistaken for a true allele of the sample, complicating the determination of a<br />

genotype or even resulting in false genotyping. Because a PCR may not be affected<br />

each time, independent amplifications of the same sample may result in differing<br />

genotyping results (e.g., see ref. 6) for the same sample (2,7,8).<br />

The degree at which these two artifacts occur is related to quantity and quality<br />

of extracted DNA amplified within a PCR (1,4,8,9). Therefore, an optimized DNA<br />

extraction, to get the highest possible amount of target DNA with the best possible<br />

quality (which means with the highest possible reduction of additional degradation<br />

during the extraction procedure), is an essential precondition for optimal reproducibility<br />

of STR genotyping in the case of samples containing ancient DNA.<br />

The protocols presented here are the results of a study with the aim to optimize the<br />

extraction of ancient DNA from historical skeletal material described by Schmerer et al.<br />

(10,11). (For further detailed <strong>info</strong>rmation on this study, also refer to ref. 12.)<br />

The protocol described in detail represents a consensus protocol developed for the<br />

extraction of DNA with a wide range in degree of degradation (cf. ref. 10) and may<br />

be used regardless of the state of DNA preservation present in the skeletal material<br />

intended for analysis. Additional <strong>info</strong>rmation on special protocols designed for three<br />

different degrees of DNA degradation is given in the table (see Note 1) and can be used<br />

to adapt the protocol in case of known DNA preservation.<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

57


58 Schmerer<br />

2. Materials<br />

To apply the protocol described in the next section, the following chemicals and<br />

solutions are required.<br />

1. 0.5 M EDTA solution (pH 8.3).<br />

2. Sterile water (Ampuwa ® , Fresenius).<br />

3. Proteinase K-solution (approx 2 g/mL, e.g. Perkin–Elmer).<br />

4. 70% phenol/chloroform/water (24251, e.g., Rotipuran, Roth, or Perkin–Elmer) or,<br />

alternatively, a 5 M solution of sodium perchlorate in sterile water.<br />

5. Chloroform (100%, e.g., Perkin–Elmer).<br />

6. 2 M sodium acetate buffer (pH 4.5, Perkin–Elmer).<br />

7. Isopropanol (abs., Merck).<br />

8. Silica solution (Glasmilk , Dianova).<br />

9. Ethanol (abs.).<br />

3. Methods<br />

1. To prevent coprocessing of possible adhering contaminations, exposed surfaces of the<br />

bone sample are removed by the use of a scalpel. Subsequently, the material is exposed to<br />

ultraviolet light for 15 min on each of the surfaces.<br />

2. According to the consistency of the material, samples are ground to a fine powder using a<br />

mixer mill (MM2, Retsch) or a mortar and pestle.<br />

3. Bone powder (0.3 g) is mixed with 1.5 mL of 0.5 M EDTA-solution (pH 8.3) in a 2-mL<br />

reaction tube (Eppendorf), vortexed vigorously, and incubated for 96 h in a shaking<br />

waterbath at 20°C and constant shaking (for decalcification of the bone material).<br />

4. Residues of the bone powder are pelleted by centrifugation at 3000g for 5 min in a 5415 C<br />

bench-top centrifuge (Eppendorf) or equivalent. The following steps are performed with the<br />

supernatants (approx 1.3 mL). These are transferred to an automated DNA extraction system<br />

(Gene Pure, Perkin–Elmer) or alternatively to a 15-mL tube (e.g., BlueMax , Falcon).<br />

5. The aqueous supernatants are mixed with 1.3 to 1.8 mL sterile water and incubated with<br />

380 to 650 µL proteinase K-solution (approx 2 g/mL) at 60°C for 1.5 h.<br />

6. Two volumes (2× supernatant volume) of 70% phenol/chloroform/water (24251) are<br />

added to the solution. The suspension is constantly shaken for 6 min at room temperature.<br />

Alternatively, phenol can be replaced by 2 mL of a 5 M sodium perchlorate-solution (ref. 13,<br />

see Notes 2 and 3).<br />

7. Using phenol—to facilitate the process—phase separation is performed at 60°C for 8 min<br />

and the phenolic layer is removed. Alternatively, the suspension can be separated by<br />

centrifugation at 4500g for 10 min. Using sodium perchlorate, no phase separation occurs.<br />

After incubation at 60°C for 8 min, the extraction mix is therefore processed according<br />

to step 8.<br />

8. The aqueous phase is mixed with 4.0 to 5.3 mL of chloroform (100%) and the resulting<br />

suspension shaken for 6 min at room temperature.<br />

9. Phase separation is performed at 60°C for 8 min and the chloroform phase is subsequently<br />

removed. Alternatively, the suspension can be separated by centrifugation at 4500g for<br />

10 min.<br />

10. DNA precipitation takes place in the presence of 64 to 120 µL of 2 M sodium acetate-buffer<br />

(pH 4.5) and 2.8 to 3.8 mL of isopropanol (abs.) at room temperature. After mixing these<br />

components for 1 min, the pH of the extraction mix should be determined. If necessary,<br />

the pH value should be adjusted to a maximum of 7.5 by further addition of sodium<br />

acetate buffer to ensure optimal precipitation of DNA (see Note 4). Subsequently 5 µL


Extraction of Ancient DNA 59<br />

Table 1<br />

Extraction Parameters<br />

DNA preservation state *<br />

Intermediate<br />

Low<br />

Parameters High degradation degradation degradation<br />

Time of EDTA incubation (96)–120 h 48 h 24 h<br />

Temperature of EDTA 20°C (20)–30°C (20)–30°C<br />

incubation<br />

Time of proteinase (60)–90 min 90 min 60 min<br />

K incubation<br />

Extraction reagent Sodium perchlorate Phenol (sodium Phenol (sodium<br />

perchlorate)<br />

perchlorate)<br />

Amount of sodium acetate 64–117 µL 64–117 µL 64–117 µL<br />

solution (2 M) to to<br />

(128–233 µL) (128–233 µL)<br />

Additional purification Wizard Prep (Wizard Prep) (Wizard Prep)<br />

* DNA preservation state was determined by PCR amplification success and reproducibility of amplification<br />

results applying the basic protocol optimization was based upon (2,5,16).<br />

The preservation state was defined as high degradation when specific products were detectable in up<br />

to 35% of the amplifications, displaying a low degree of reproducibility (12%) in typing results, and low<br />

degradation when specific products were detected in 70 to 90% of the amplificates, with a reproducibility<br />

of 86–91%.<br />

of silica-solution (Glasmilk , Dianova) is added and the mix is shaken again for 10 min<br />

at room temperature.<br />

11. Using a DNA extraction device (Gene Pure, Perkin–Elmer), the mix is removed by pressure<br />

and the DNA–silica complex is collected on a filter membrane. Alternatively, the mixture<br />

can be separated by centrifugation at 4500g for 10 min.<br />

12. The DNA–silica complex is washed with 2.8 to 3.8 mL of ethanol (abs.) for 5 min at<br />

room temperature. Then, the alcohol is removed by renewed filtration or centrifugation.<br />

Co-extracted salts may be removed using ethane (80Y).<br />

13. Using an automated DNA extractor, the silica-bound DNA is manually removed from<br />

the filter membrane with 500 µL of ethanol (abs.) and transferred to a 2-mL reaction<br />

tube for further processing.<br />

14. The DNA-silica complex is pelleted by centrifugation for 4500g for 4 min (Centrifuge<br />

5804, Eppendorf) and the ethanol is discarded.<br />

15. The resulting pellet is air-dried for approx 30 min at room temperature, redissolved in 50 µL<br />

of sterile water (Ampuwa ® , Fresenius), thermally eluted at 50°C for 5 min at constant<br />

shaking (thermal shaker 5437, Eppendorf), and stored at –20°C for further processing.<br />

4. Notes<br />

1. In case the state of DNA preservation of the material worked on is known, please change<br />

the parameters concerned according to Table 1.<br />

2. The use of sodium perchlorate may result in an inhibition of DNA polymerase during the<br />

amplification process and therefore requires a further cleaning of the extracted DNA. For<br />

this purpose, the Wizard PCR Preps purification system (Promega) following a modified<br />

protocol (12,14) may be used.


60 Schmerer<br />

3. The additional purification of the extract necessary when using sodium perchlorate results<br />

in loss of DNA. The use of sodium perchlorate, therefore, might be beneficial if brownish<br />

color of the bone meal or the extract indicates the presence of co-extracted polymeraseinhibiting<br />

substances like humic acids (15). In this case, an additional cleaning step would<br />

always be necessary and the use of sodium perchlorate instead of phenol (13) would not<br />

result in further loss of DNA.<br />

4. Make sure to add the sodium-acetate buffer first, and to adjust the pH before adding<br />

isopropanol to avoid co-precipitation of salts.<br />

5. Safety Note<br />

Phenol and chloroform are aggressive organic solvents (please notice the safety<br />

data sheets provided by the manufacturers) and volatile even at room temperature,<br />

especially at the elevated temperature recommended to facilitate phase separation in<br />

steps 7 and 9. Avoid skin contact and inhalation. The use of a vented chemical hood<br />

is strongly recommended.<br />

References<br />

1. Kimpton, C., Fisher, D., Watson, S., Adams, M., Urquhard, A., Lygo, J., et al. (1994)<br />

Evaluation of an automated DNA profiling system employing multiplex amplification of<br />

four tetrameric STR loci. Int. J. Leg. Med. 106, 302–311.<br />

2. Schmerer, W. M., Hummel, S., and Herrmann, B. (1997) Reproduzierbarkeit von aDNAtyping.<br />

Anthrop. Anz. 55, 199–206.<br />

3. Weber, J. L. and May, P. E. (1989) Abundant class of human polymorphisms which can be<br />

typed using the polymerase chain reaction. Am. J. Hum. Genet. 44, 388–396.<br />

4. Murray, V., Monchawin, C., and England, P. R. (1993) The determination of the Sequences<br />

present in the shadow bands of a dinucleotide repeat PCR. Nucleic Acids Res. 21,<br />

2395–2398.<br />

5. Schmerer, W. M. (1996) Reproduzierbarkeit von Mikrosatelliten-DNA-Amplifikation und<br />

Alleldetermination aus bodengelagertem Skelettmaterial. Unpublished diploma-thesis,<br />

Göttingen.<br />

6. Ramos, M. D., Lalueza, C., Girbau, E., Perez-Perez, A., Quevedo, S., Turbon, D., et al.<br />

(1995) Amplifying dinucleotide microsatellite loci from bone and tooth samples of up to<br />

5000 years of age: More inconsistency than usefulness. Hum. Genet. 96, 205–212.<br />

7. Schultes, T., Hummel, S., and Herrmann, B. (1997) Recognizing and overcoming inconsistencies<br />

in microsatellite typing of ancient DNA samples. Ancient <strong>Bio</strong>mol. 1, 227–233.<br />

8. Ivanov, P. L. and Isaenko, M. V. (1999) Identification of human decomposed remains<br />

using the STR systems: effect on typing results. In: Proceedings of the Second European<br />

Symposium on Human Identification 1998. Promega Corporation, Innsbruck, Austria.<br />

9. Lygo, J. E., <strong>John</strong>son, P. E., Holaway, D. J., Woodroffe, S., Whitaker, J. P., Clayton, T. M.,<br />

et al. (1994) The validation of short tandem repeat (STR) loci for use in forensic casework.<br />

Int. J. Leg. Med. 107, 77–89.<br />

10. Schmerer, W. M., Hummel, S., and Herrmann, B. (1999) Optimized DNA extraction to<br />

improve reproducibility of short tandem repeat genotyping with highly degraded DNA as<br />

target. Electrophoresis 20, 1712–1716.<br />

11. Schmerer, W. M., Hummel, S., and Herrmann, B. (2000) STR-genotyping of archaeological<br />

human bone: Experimental design to improve reproducibility by optimisation of DNA<br />

extraction. Anthrop. Anz. 58, 29–35.<br />

12. Schmerer, W. M. (2000) Optimierung der STR-Genotypenanalyse an Extrakten alter DNA<br />

aus bodengelagertem menschlichen Skelettmaterial. Cuvillier Verlag, Göttingen.


Extraction of Ancient DNA 61<br />

13. <strong>John</strong>s, M. B. and Paulus-Thomas, J. E. (1989) Purification of human genomic DNA from<br />

whole blood using sodium perchlorate in place of phenol. Anal. <strong>Bio</strong>chem. 180, 276–278.<br />

14. Lassen, C. (1998) Molekulare Geschlechtsdetermination der Traufkinder des Gräberfeldes<br />

Aegerten (Schweiz). Cuvillier Verlag, Göttingen.<br />

15. Cooper, A. (1992) Removal of colourings, inhibitors of PCR, and the carrier effect of PCR<br />

contamination from ancient DNA samples. Ancient DNA Newsletter 1, 31–33.<br />

16. Burger, J., Hummel, S., and Herrmann, B. (1997) Nachweis von DNA-Einzelkopiesequenzen<br />

aus prähistorischen Zähen. Liegemilieu als Faktor für den Erhalt von DNA. Anthrop.<br />

Anz. 55, 193–198.


62 Schmerer


DNA Extraction from Plasma and Serum 63<br />

16<br />

DNA Extraction from Plasma and Serum<br />

David Stirling<br />

1. Introduction<br />

There are occasions where the only materiel available on a patient is stored plasma or<br />

serum samples. In normal individuals, the amount of DNA in these samples is very low<br />

but sufficient to serve as template for PCRs. Moreover, increased amounts of circulating<br />

DNA have been found in a variety of disorders, including cancer, autoimmune disease,<br />

and infection. Additionally, small amounts of fetal DNA have been detected in maternal<br />

plasma/serum during gestation. We have used the following protocol to successfully<br />

genotype archival plasma samples.<br />

2. Materials<br />

1. 10X SDS /Protein K: (Lauryl sulphate [SDS] 10 g/100 mL, Proteinase K 5 mg/mL).<br />

2. TE (Tris EDTA) buffer: 10 mM Tris-HCl, 1 mM EDTA, pH 8.0.<br />

3. Phenolchloroform (11 v/v).<br />

4. Glycogen (10 mg/mL).<br />

5. 7.5 M Ammonium acetate.<br />

6. 100% ethanol.<br />

7. 70% ethanol.<br />

3. Method<br />

1. Place 1.5 mL of serum or plasma into a 15-mL centrifuge tube.<br />

2. Add 1.5 mL of 1X SDS proteinase K solution in the tube containing the serum and mix<br />

well.<br />

3. Digest overnight at 55°C in water bath.<br />

4. Add 3 mL of phenol/chloroform solution.<br />

5. Vortex 30 s and centrifuge for 10 min at 1000g using a swing-out rotor.<br />

6. Transfer aqueous layer to fresh tube and repeat steps 4 and 5.<br />

7. Transfer aqueous layer to fresh tube and add 5 µL of glycogen (10 mg/L), 1 mL of 7.5 M<br />

ammonium acetate, and 8 mL of 100% ethanol.<br />

8. Mix by inverting and centrifuge at 2500g for 40 min.<br />

9. Carefully remove supernatant and wash pellet in 10 mL of 70% ethanol.<br />

10. Centrifuge at 2500g for 10 min. Carefully remove last traces of ethanol, and allow to air<br />

dry for 10 min before redissolving in 100 µL of TE.<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

63


64 Stirling


Detection of Nucleic Acids 65<br />

17<br />

Technical Notes for the Detection of Nucleic Acids<br />

<strong>John</strong> M. S. <strong>Bartlett</strong><br />

1. Introduction<br />

In following any polymerase chain reaction (PCR)-based method, it is usual to<br />

identify the products of the reaction by some form of detection system. The majority<br />

of these still rely on size- and charge-based separation systems, although for some<br />

quantitative PCR applications, either direct measurement of fluorescence or indirect<br />

Enzyme-linked immunosorbent assay-based systems can be used. In this chapter, we<br />

summarize some of the most common methods for detection of nucleic acids as a<br />

handy reference for those seeking to validate their PCR reactions. Although there are<br />

many variants of these techniques, we have confined our reporting to methods of which<br />

we have direct experience and which offer broad applicability. Fluorescence detection<br />

of quantitative real-time PCR requires specialist equipment, and we have therefore<br />

omitted this from our discussions at present.<br />

2. Gel Electrophoresis<br />

By far the most common procedure for the analysis of nucleic acids is gel electrophoresis.<br />

This is a highly flexible approach that provides <strong>info</strong>rmation on the size of<br />

the DNA molecule and under certain conditions can be used to discriminate different<br />

sequences of the same size. Electrophoresis can also be used to separate and purify<br />

nucleic acid fragments, to quantify allelic imbalances, etc. For DNA electrophoresis,<br />

the most common supports used are agarose and polyacrylamide. These are highly<br />

flexible because varying the concentration, and for polyacrylamide, the degree of<br />

crosslinking, can markedly alter the size range which can be discriminated. In general,<br />

polyacrylamide gels are more useful for separating smaller fragments of DNA (under<br />

300–500 base pairs) and for applications where high resolution is required (such as<br />

analysis of microsatellites) because they are capable of resolving size differences of<br />

as little as 1 bp. Polyacrylamide gels can be run faster and at higher temperature<br />

than agarose gels. However, acrylamide is a neurotoxin and presents a safety hazard.<br />

Most laboratories now eschew the process of producing their own acrylamide solutions,<br />

relying instead on commercially prepared materials to circumvent this hazard.<br />

Polyacrylamide gels are also more difficult to pour and handle than agarose gels.<br />

Furthermore, using polyacrylamide gels requires, in general, the use of labeled<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

65


66 <strong>Bartlett</strong><br />

Table 1<br />

Recommended Agarose Concentrations<br />

for DNA Electrophoresis<br />

Percentage agarose (w/v)<br />

Molecular weight range<br />

0.6% 1000 –20,000 bp<br />

long PCR<br />

1.0% 1500 –5000 bp<br />

2% 1100 –2000 bp<br />

3% 1110 –500 bp<br />

nucleotides, although silver staining may be applied to these gels the fragility of the<br />

gel often precludes this.<br />

Agarose gels, however, are more robust and easy to prepare. Although resolution is<br />

poorer, some modern forms of agarose claim separations that rival that of acrylamide<br />

gels. The major strength of agarose-based gels is the greater range of separation.<br />

Conventional agarose electrophoresis can separate DNAs from 200 to 50,000 bp which<br />

is more than adequate for PCR-based systems. Adaptations of agarose electrophoresis,<br />

for example, those using pulsed electric fields, can be used to separate DNA fragments<br />

of up to 10 Mbp.<br />

2.1. Selecting Conditions for Agarose Gel Electrophoresis<br />

Those interested in understanding the electrophysical properties governing migration<br />

of DNA in agarose supports can refer to a number of molecular biology texts that<br />

cover these areas or alternatively visit the web (1) for a useful guide to electrophoresis<br />

with agarose gels.<br />

2.1.1. Agarose Concentrations<br />

Table 1 outlines recommended agarose concentrations for gel electrophoresis of<br />

DNA. Note that modern specialist formulations of agarose may require alterations to<br />

this table (e.g., Nusieve, etc.).<br />

2.1.2. Buffers<br />

Two buffering systems are commonly used for agarose gel electrophoresis of DNA;<br />

of these, Tris acetate EDTA buffer (TAE) is more widely accepted because it facilitates<br />

recovery of material from agarose. However, it has a relatively low buffering capacity,<br />

and recirculation of buffer may be required over long electrophoresis runs (4–6 h).<br />

Tris borate EDTA (TBE) is preferred for small molecules and longer electrophoresis<br />

times because of its higher buffering capacity. We have used both with good results.<br />

Independent of the buffer selected, attention should be paid to the depth of buffer. In<br />

most systems 3 to 5 mm of buffer should cover the gel. Insufficient buffer may allow the<br />

gel to dry out during the run, whereas excessive buffer will reduce the current through<br />

the agarose support, promote heating and decrease DNA motility. Electrophoresis is<br />

performed applying a voltage of between 1 to 5 V/cm (where cm is the distance in<br />

centimeters between the electrodes in the gel tank).


Detection of Nucleic Acids 67<br />

2.1.3. Loading Dyes<br />

Loading dyes serve two purposes: They are usually dense and promote the settling<br />

of the DNA to the base of the well. They usually contain dyes that migrate with the<br />

DNA and can be visualized to monitor the process of electrophoresis. Care should<br />

be taken, however, because the common dyes used (bromophenol blue and xylene<br />

cyanol) have different motilities in different agarose products, with differing buffers<br />

and differing gel percentages (see ref. 3).<br />

2.1.4. DNA Dyes: Before or After?<br />

Most dyes (ethidium bromide, SYBR green, etc.) that are used to stain DNA do<br />

so by intercalating into the DNA sequence. As such they, of necessity, alter both the<br />

structure and motility of the DNA. Although this can give spurious results, for many<br />

applications (such as confirmation of PCR products before cloning or sequencing) this<br />

is not a major issue. In these cases, the dye is often added to the loading buffer or gel<br />

before electrophoresis. Where an accurate determination of product size is required,<br />

such as in randomly amplified polymorphic DNA (RAPD) or allelotyping, products<br />

should be separated in the absence of dye and the gel stained thereafter.<br />

2.1.5. Recovery of DNA from Agarose Gels<br />

Having identified the product on agarose gels, often there is a requirement to<br />

sequence or otherwise analyze the product. Recovering DNA from agarose gels is a<br />

simple procedure that can be attempted in many ways (1). It is important, however, that<br />

the decision regarding recovery of DNA is taken before electrophoresis is attempted<br />

as the choice of agarose is probably the most crucial factor in determining DNA yield<br />

after extraction from agarose. Protocols for recovery of DNA from agarose can be<br />

accessed from the web (see ref. 1 for a good range of methods).<br />

2.2. Polyacrylamide Gel Electrophoresis (PAGE) of DNA<br />

Casting and running polyacrylamide gels is undoubtedly more complex and problematic<br />

than using agarose electrophoresis. Almost universally, polyacrylamide gels<br />

are supported between glass plates and must be polymerized by a chemically catalyzed<br />

reaction with the inherent problems caused by failed reagents from time to time.<br />

However, PAGE has distinct advantages in terms of resolution, capacity, and purity of<br />

DNA bands. The resolving power of PAGE is such that molecules of DNA with lengths<br />

differing by as little as 0.2% (2 bases/Kb) may be separated (although some novel<br />

agarose preparations can approach this for smaller fragments). PAGE has a higher<br />

capacity for DNA than agarose, with up to 10 µg of DNA per lane being applied<br />

without compromising results. DNA recovered from PAGE is extremely pure and has<br />

been used for microinjection into mouse embryos (2).<br />

PAGE is equally applicable to separation of double-stranded or single-stranded<br />

DNA (under denaturing conditions); however, the size range of fragments that can<br />

be separated is more restrictive than for agarose because of the fragile nature of low<br />

percentage acrylamide gels (see Table 2).<br />

PAGE gels are crosslinked using a varying ratio of bis-acrylamide in the monomeric<br />

solution; generally, the ratio of acrylamide to bis-acrylamide is around 301.


68 <strong>Bartlett</strong><br />

Table 2<br />

Separation of Double-Stranded DNA:<br />

What Percentage of Acrylamide?<br />

Acrylamide (% w/v)<br />

DNA size range<br />

13.5 1000 –2000 bp<br />

15.0 1180 –500 bp<br />

18.0 1160 – 400 bp<br />

12.0 1140 –200 bp<br />

15.0 1125 –150 bp<br />

20.0 1116 –100 bp<br />

2.2.1. Nondenaturing PAGE Conditions<br />

Although both TAE and TBE can be used for PAGE, we recommend use of TBE<br />

because this generally provides sharper bands on low-percentage gels. Running gels in<br />

1× TBE under low voltages (1–8 V/cm) will prevent denaturation of DNA caused by<br />

heating of the gel. Although the electrophoretic mobility of double-stranded DNA<br />

is inversely proportional to the log(fragment length), this relationship can be altered<br />

by both GC content and sequence. For this reason, PAGE is not a reliable means of<br />

sizing DNA fragments.<br />

2.2.2. Denaturing PAGE<br />

Denaturing page gels contain a denaturing agent (usually 6–8 M urea) that inhibits<br />

base pairing in nucleic acids. DNA fragments are loaded after a brief heat denaturation<br />

and electrophoresis at high voltages (ca 20 V/cm) ensures that high temperatures are<br />

maintained throughout the electrophoretic process. Denaturing gels are most commonly<br />

used to analyze sequencing reactions and to recover labeled DNA probes. As one<br />

would predict from their common use in sequence determination, the motility of DNA<br />

fragment under denaturing conditions is almost totally unaffected by sequence and<br />

base composition.<br />

2.2.3. Recovery of DNA from Polyacrylamide Gels<br />

In our experience, recovery of DNA from polyacrylamide gels is vastly simpler than<br />

recovery from agarose, largely because of the fact that much larger quantities of DNA<br />

are loaded onto the gel. Elution overnight by diffusion can recover 70 to 80% of the<br />

nucleic acid in Maxim–Gilbert’s solution.<br />

2.2.4. Radioisotopic Detection and Quantitation<br />

One of the major advantages of polyacrylamide gels is the ability to dry them<br />

onto supports and produce autoradiographic images. These can either be analyzed by<br />

densitometry or used as a template to allow direct counting of the bands of interest.<br />

Although the latter approach is time consuming, it provides a highly quantitative<br />

analysis of DNA species in (for example) quantitative PCR (3).


Detection of Nucleic Acids 69<br />

2.2.5. Nonradioisotopic Detection and Quantitation<br />

The increasing trend in modern molecular applications is away from the use of large<br />

amounts of radioactivity and to substitute biotinylated or digoxigenin-labeled DNA<br />

for 32-P- or 33-P-labeled nucleotides. More recently, directly fluorescently labeled<br />

nucleotides have been released which, with the appropriate systems, can also aid<br />

detection of nucleic acid. These approaches are largely applicable in northern and<br />

southern blotting procedures, and although these can be applied to sequencing (blotting<br />

of sequences) or other DNA gel procedures, they are time consuming and introduce an<br />

additional potential source of error. Other methods, such as silver staining, are more<br />

attractive because they do not require blotting, but they are generally less sensitive than<br />

radioisotopic procedures. With this proviso, they provide a useful and rapid means of<br />

detecting nucleic acids in situ, before proceeding with experiments.<br />

2.3. Summary<br />

It is of course impossible to detail all possible systems for detection of DNA in a<br />

simple chapter such as this one; however, we have included below examples of the<br />

major approaches to detection of nucleic acids that can be adapted to suit most needs.<br />

This is very much a general guide for broad application and other methods can be<br />

sourced from the references included below.<br />

3. Materials<br />

3.1. Agarose Gel Electrophoresis<br />

1. Low melting point agarose (Sigma).<br />

2. 50× TAE: 242 g of Trizma Base (Sigma) and 100 mL of 0.5 M EDTA (pH 8.0) are<br />

dissolved in 800 mL distilled water and autoclaved. Then, 57.1 mL of glacial acetic acid<br />

is added; make up to 1 L.<br />

3. Gel loading buffer: 0.25% Bromophenol blue (w/v), 0.25% xylene cyanol FF (w/v), and<br />

30% glycerol (v/v) in distilled water (see Note 1).<br />

4. Molecular weight markers (e.g., 100-bp ladder; Gibco).<br />

5. Gel apparatus and power pack.<br />

6. Ethidium bromide (10 mg/mL) in water (see Note 2).<br />

7. Ultraviolet light box and camera.<br />

8. Activated charcoal.<br />

3.1.1 Recovery of DNA from Agarose Gels<br />

1. Scalpel blades.<br />

2. β-agarose and buffer (Calbiochem, UK).<br />

3. Phenolchloroform 11 mixture (Sigma, UK).<br />

4. Ammonium acetate (7.5 M).<br />

5. 100% Ethanol.<br />

3.1.2. PAGE<br />

3.1.2.1. NONDENATURING PAGE<br />

1. 40% Acrylamide (38 g of acrylamide, 2 g of bis-acrylamide, <strong>Bio</strong>–Rad, UK) in distilled<br />

water.


70 <strong>Bartlett</strong><br />

2. 5× TBE: 54 g of Tris-base, 27.5 g of boric acid, 20 mL of 0.5 M EDTA (pH 8.0) dissolved<br />

in 800 mL distilled water; make up 1 L.<br />

3. 10% w/v ammonium persulphate (make fresh for each use).<br />

4. TEMED (<strong>Bio</strong>–Rad).<br />

5. Gel loading buffer: 0.25% Bromophenol blue (w/v), 0.25% xylene cyanol FF (w/v), and<br />

30% glycerol (v/v) in distilled water (see Note 1).<br />

6. Molecular weight markers (e.g., 100-bp ladder; Gibco).<br />

7. Vertical gel electrophoresis system, spacers (1.0 mm; e.g., <strong>Bio</strong>–Rad Protean II).<br />

3.1.2.2. DENATURING PAGE<br />

1. 40% Acrylamide (38 g of acrylamide, 2 g of bis-acrylamide, <strong>Bio</strong>–Rad, UK) in distilled<br />

water.<br />

2. 5× TBE: 54 g of Tris-base, 27.5 g of boric acid, 20 mL of 0.5 M EDTA (pH 8.0) dissolved<br />

in 800 mL of distilled water; make up to 1 L.<br />

3. Urea (Sigma, UK).<br />

4. 10% w/v ammonium persulphate (make fresh for each use).<br />

5. Temed (<strong>Bio</strong>–Rad).<br />

6. Radioactively labeled molecular weight markers: We have used Hinf1-digested plasmids<br />

labeled with Klenow enzyme (see Note 3).<br />

7. Gel loading buffer: 0.25% Bromophenol blue (w/v), 0.25% xylene cyanol FF (w/v), and<br />

30% glycerol (v/v) in distilled water (see Note 1).<br />

8. Vertical gel electrophoresis system, spacers (0.6 mm).<br />

9. Gel fixative: Prepare 2 L of 10% acetic acid and 10% methanol in distilled water.<br />

10. Whatmann 3MM Paper.<br />

11. Gel dryer.<br />

12. Autoradiographic film and cassettes.<br />

13. Developer and fixative for above.<br />

3.1.3. Recovery of DNA from Acrylamide Gels<br />

1. Maxam & Gilberts solution: 0.5 M ammonium acetate, 0.1% sodium dodecyl sulphate,<br />

and 1 mM EDTA; store at –20°C.<br />

2. Transfer RNA (tRNA) 10 mg/mL and store at –20°C (optional).<br />

3. Absolute ethanol (–20°C).<br />

4. 70% ethanol (–20°C).<br />

4. Methods<br />

4.1. Agarose Gel Electrophoresis<br />

1. Prepare a clean flat bed electrophoresis tray, with sealed ends (see Note 4) and a comb<br />

with the appropriate number of samples. Fill tank with 1× TAE.<br />

2. Dissolve 3 g of low melting point agar in 100 mL of 1× TAE buffer.<br />

3. Weigh the flask and note weight before microwaving.<br />

4. Microwave until clear, take care not to over boil (see Note 5).<br />

5. Allow to cool slightly, weigh, and add sufficient distilled water to make up to premicrowave<br />

weight.<br />

6. Cool agarose to approx 50°C before pouring into gel support. Insert comb 0.5 to 1.0 mm<br />

above the base of the tank, add agarose and allow to set (30 to 60 min).<br />

7. Remove buffer dams or tape and place gel in electrophoresis tank. Add sufficient buffer<br />

to cover gel to approx 1 mm.


Detection of Nucleic Acids 71<br />

8. Mix DNA with gel loading buffer by adding 11 DNA solution to gel loading buffer.<br />

Slowly add approx 5 to 10 µL sample per well. Add 1 µg of DNA molecular weight marker<br />

to 4 µL of water and 5 µL of gel loading buffer to one lane (see Note 6).<br />

9. Close the lid of the tank to generate an electrical circuit (see Note 7) and run at a voltage<br />

of 1 to 5 volts/cm. Check bubbles are arising from the anode and that the dye is migrating<br />

into the gel.<br />

10. When migration of markers is complete, remove gel and visualize under ultraviolet light<br />

and photograph.<br />

11. Use intensity of bands to estimate DNA concentration; densitometric scanning of the<br />

photograph can also be used as appropriate.<br />

4.1.1. Staining of Agarose or Acrylamide Gels<br />

1. Remove the gel from the electrophoresis tank and stain in 0.5 to 1 µg/mL ethidium<br />

bromide in sufficient electrophoresis buffer to cover the gel for about 10 to 20 min at<br />

room temperature.<br />

2. Destain in distilled water for 2 × 20 min.<br />

3. Visualize gel (see Note 7).<br />

4.1.2. Decontamination of Ethidium Bromide Solutions<br />

Both electrophoresis buffers from ethidium bromide containing gels and staining/<br />

destaining of gels should be decontaminated before disposal (see Note 2).<br />

1. Add 1 g/L activated charcoal.<br />

2. Stir for 30 to 60 min at room temperature.<br />

3. Remove charcoal by filtration and discard solution.<br />

4. Place charcoal and agarose gel into solid waste for incineration (see Note 8).<br />

4.1.3. Recovery of DNA from Agarose Gels<br />

Recovery of DNA from agarose gels produces highly variable yields dependent on<br />

fragment length. Long fragments (5–10 kb and above) are prone to shearing if vortexed.<br />

We have described here the agarose method that is most widely applicable, in our<br />

hands; however, postrecovery purification with phenol chloroform greatly improves<br />

the purity of the DNA and is recommended where postrecovery enzyme modifications<br />

are planned. The use of TAE buffer is recommended and care should be taken not<br />

to overload the gel.<br />

1. After running the gel, visualize the band of interest under ultraviolet light and excise using<br />

a scalpel (see Note 9). Remove the band as cleanly as possible; the lower the amount of<br />

agar present in the gel slice, the better the recovery.<br />

2. Transfer up to 200 mg of agarose gel slice to a microcentrifuge tube and melt at 75°C<br />

(5–10 min. see Note 10) then cool to 45°C.<br />

3. Estimate volume of melted gel and add 2% 50× agarose buffer and mix gently (see<br />

Note 11).<br />

4. Add β-agarose (add 3–4 units per 100 mg of a 1% agarose gel), mix gently (see Note 11),<br />

and incubate at 45°C for 3 to 4 h (see Note 12).<br />

5. Add an equal volume of phenolchloroform and mix by gentle inversion 10 to 20 times.<br />

6. Centrifuge at 3000g for 5 to 10 min.


72 <strong>Bartlett</strong><br />

7. Remove upper aqueous layer to a new microfuge tube and add 20 µL of 7.5 M ammonium<br />

acetate and 2 volumes of ice cold 100% ethanol. Mix by gentle inversion 10 to 20 times<br />

and let stand for 30 min (see Note 13).<br />

8. Centrifuge at 15,000g for 10 min and discard supernatant.<br />

9. Allow pellet to air dry and resuspend in distilled water.<br />

4.2. PAGE<br />

Many applications using PAGE are based on radiolabeling of DNA; however, we<br />

have successfully used this system also for silver-stained and ethidium bromide-stained<br />

gels.<br />

4.2.1. Nondenaturing PAGE<br />

1. Prepare a 6% (see Note 14) acrylamide gel by mixing 7.5 mL of 40% acrylamide,<br />

10 mL of 5×TBE, and 32.5 mL of distilled water (50 mL total, sufficient for 2 × 20 cm<br />

protean II gels).<br />

2. The protean II system is supplied with a gel-pouring apparatus, the vertical clamps seal the<br />

sides of the gel, and a rubber cushion seals the base. Add 25 µL of TEMED and 250 µL of<br />

fresh ammonium persulphate solution to the acrylamide and pour gels immediately (see<br />

Note 15). Insert combs and leave to polymerize for 30 to 60 min.<br />

3. Remove combs and wash wells with 1× TBE (see Note 16). Prepare electrophoresis<br />

equipment and add 1× TBE to top and bottom buffer chambers. Cool electrophoresis<br />

equipment by passing water through central core (see Note 17).<br />

4. Mix DNA with gel loading buffer by adding 11 DNA solution to gel loading buffer.<br />

Slowly add approx 5 to 10 µL of sample per well. Add 1 µg of DNA molecular weight<br />

marker to 4 µL of water and 5 µL of gel loading buffer to one lane (see Note 6).<br />

5. Electrophorese at 30 mA for 2 to 3 h.<br />

6. Carefully remove the top gel plate and thoroughly wet gloves and gel with 1× TBE.<br />

Gently remove gel onto a prewetted support prior to visualization under ultraviolet light<br />

(see Note 18).<br />

4.2.2. Denaturing PAGE<br />

Although nondenaturing PAGE is applicable to both unlabeled and labeled DNA,<br />

the fragile nature of the thin gels commonly used for denaturing PAGE makes radioactive<br />

labeling more common. We have, however, successfully used silver staining to<br />

visualize DNA on PAGE (see ref. 1). We have, however, described here conventional<br />

radioactive detection of single-stranded DNA.<br />

1. To make a 6% denaturing gel: weigh 42 g of urea and add 15 mL of 40% acrylamide,<br />

20 mL of 5× TBE, and 30 mL of distilled water. Stir at room temperature until urea<br />

dissolves completely.<br />

2. Prepare vertical gel electrophoresis system for gel pouring, ensuring that the plates are<br />

clean (see Note 19), and seal with sleek tape.<br />

2. Add 80 µL of TEMED and 800 µL of fresh ammonium persulphate solution to the<br />

acrylamide and pour gel immediately (see Note 15), tapping vigorously during pouring to<br />

free any air bubbles. Insert comb and leave to polymerize for 30 to 60 min.<br />

3. Place the gel in a vertical electrophoresis tank and fill the upper and lower chambers<br />

with 1× TBE.<br />

4. Mix DNA with gel loading buffer by adding 11 DNA solution to gel loading buffer.<br />

Slowly add approx 1 to 3 µL of sample per well. Heat samples to 85°C for 5 min and cool


Detection of Nucleic Acids 73<br />

on ice before loading. Add 1 µg of radioactively labeled DNA molecular weight marker to<br />

1 µL of gel loading buffer in one lane.<br />

5. Electrophorese at 30 mA for 2 to 3 h or at constant temperature (see Note 20).<br />

6. Remove gel from the electrophoresis tank and discard radioactive buffer.<br />

7. Place gel plates on tissue paper and insert a scalpel between the back plate and spacers.<br />

Remove the backplate trying to ensure that the gel remains on the front plate (see Note<br />

21).<br />

8. Immerse gel, on plate, in gel fixation solution and fix for 15 to 30 min. Carefully remove<br />

the gel, on the plate, from the fixative, drain, and cover with a dry piece of Whatman<br />

3MM filter paper.<br />

9. Either invert the gel and carefully lift the glass plate away from the paper or lift the paper<br />

gently from the glass plate. The gel should stick to the paper.<br />

10. Dry on a gel dryer for 60 min at 80°C under vacuum (see Note 22).<br />

11. Cover gel with clingfilm and place autoradiography film directly over the gel and expose<br />

at –70°C (see Note 23).<br />

4.2.3. Recovery of DNA from Acrylamide Gels<br />

Recovery of DNA from acrylamide is a relatively simple procedure because of the<br />

higher amounts of DNA that can be loaded on acrylamide gels, with yields of around<br />

50 to 60% sufficient DNA recovered for most applications. We have only applied this<br />

technique to unfixed gels stained with ethidium bromide.<br />

1. After running the gel, visualize the band of interest under ultraviolet light and excise using<br />

a scalpel (see Note 9). Remove the band as cleanly as possible; the lower the amount of<br />

acrylamide present in the gel slice, the better the recovery.<br />

2. Transfer to a microcentrifuge tube and add 400 µL of Maxam & Gilberts solution.<br />

2. Vortex and incubate overnight at 37°C.<br />

3. Spin down the gel by pulse centrifugation for 30 s at 5000g in a bench-top centrifuge.<br />

4. Remove supernatant to a separate Eppendorf tube (approx 350 µL).<br />

5. Add 1 µL of 10 mg/mL tRNA (optional, omit if required).<br />

6. Add 2.5 volumes of ice-cold 100% ethanol and mix gently.<br />

7. Stand for 20 min and centrifuge for 10 min at 15,000g.<br />

8. Decant supernatant and add 300 µL of 70% ethanol, mix gently.<br />

9. Centrifuge for 10 min at 15,000g.<br />

10. Decant supernatant and air dry pellet.<br />

11. Resuspend in 20 µL of distilled water.<br />

5. Notes<br />

1. For some applications, lower concentrations of dyes can give better results, for example,<br />

when the DNA fragment runs close to one of the dyes. In this case, dilute loading buffer<br />

15 in 30% glycerol.<br />

2. Ethidium bromide is a carcinogen, teratogen, and mutagen. Handle with respect.<br />

3. Digesting plasmids with enzymes such as HinfI leaves 5′ overhangs that can be labeled<br />

with Klenow using the following reaction: 1 µg of digested plasmid in 5 µL of distilled<br />

water, 2 µL of 10 × repair buffer supplied with Klenow); 1 µL of a mixture containing<br />

2 mM each of dGTP, dCTP, and dTTP; 2 µL of 35 S-dATP (approx 0.5 µCi), and 1 µL of<br />

Klenow fragment of DNA polymerase (5–10 units) in a total volume of 20 µL. Incubate<br />

for 30 min at room temperature and stop reaction by adding 2 µL of 0.25 M EDTA. Store<br />

at –20°C and use 0.5 to 1 µL per gel.


74 <strong>Bartlett</strong><br />

4. Many modern electrophoresis tanks are equipped with seals or with buffer dams to allow<br />

pouring of the gel or, alternatively, one can tape over the end of the tray with autoclave<br />

tape. Ensure gels are poured on a level surface.<br />

5. Undissolved agarose appears as small lens-shaped flecks in the solution; reheat until the<br />

solution is clear. Take care not to superheat the solution because it may boil over.<br />

6. Modern electrophoresis equipment will not allow circuits to be completed until the tank is<br />

sealed. If you have an older tank without this safety feature, discard it!<br />

7. If the bands appear faint, with a dark background the gel is understained, return to buffer<br />

containing ethidium bromide and repeat. If the bands appear faint with a bright background<br />

the gel is overstained, destain in excess electrophoresis buffer (without ethidium bromide)<br />

for 10 to 20 min and revisualize.<br />

8. Standard incineration is sufficient to decontaminate ethidium bromide.<br />

9. Ultraviolet light produces strand breaks in DNA. Even exposure for short periods (30 s)<br />

can compromise DNA recovery. Ensure recovery of bands as quickly as possible. Where<br />

multiple bands are to be recovered, we have blocked the ultraviolet from parts of the gel<br />

using cardboard under the gel support. Some protocols recommend the addition of 1 mM<br />

guanosine or cytidine to the gel and electrophoresis buffers to protect DNA at the<br />

transillumination stage.<br />

10. Ensure the agarose is completely melted to optimize digestion.<br />

11. Vortexing can shear large DNA fragments, take care.<br />

12. Although recommended incubation times are shorter, we have found extending the<br />

incubation time can improve the yield and ensure complete digestion of the agarose.<br />

13. If the expected DNA concentration is less than 50 ng/µL, we recommend the addition of<br />

5 µg of tRNA to enhance precipitation.<br />

14. We have successfully worked with 4 to 20% acrylamide gels for various applications. Use<br />

of low concentration acrylamide is only possible in our experience where supports for the<br />

gels (Whatman paper) is provided and this limits the use to radioactive applications.<br />

15. Polymerization will commence as soon as these catalysts are added and is temperature and<br />

time dependent, work quickly but carefully.<br />

16. Wells will frequently contain small amounts of acrylamide solution that has notolymerisind<br />

because of contact with air (oxygen inhibits polymerization). Washing the wells prevents<br />

this from sinking to the base of the well and polymerizing giving uneven well depths.<br />

17. The protean system is cooled by circulation of water through a central core. This is helpful<br />

when rapid results are required, but not essential.<br />

18. Acrylamide gels, especially low percentage gels, are extremely fragile and tear easily, and<br />

handling them is an art learned through practice (and much frustration!). Siliconization<br />

of glass plates can aid removal of gels, but in our experience, providing the plates are<br />

clean separation is relatively easy. If your gel sticks to the glass, clean them thoroughly<br />

before use.<br />

19. A common cause of tearing of gels and problems with pouring is dirty glass plates. Before<br />

use, scrub glass plates with concentrated detergent, rinse extensively with tap water and<br />

distilled water, and then with methylated spirits. Dry thoroughly before use. Clean spacers<br />

and combs by wiping with methylated spirits.<br />

20. Modern powerpacks provide the option of running gels at constant temperature (usually<br />

65°C), and we have found this to provide more consistent results.<br />

21. The use of a silicon solution on the backplate after washing can facilitate separation. Be<br />

pragmatic: If the gel sticks to the back plate remove the front plate. When silver staining<br />

is to be used, stain the gel in situ on the plates.<br />

22. If the gel cracks into a “crazy paving” type pattern while drying, this is indicative of poor<br />

fixation. Extend fixation period. This approach can be used with thicker gels (1–1.5 mm),


Detection of Nucleic Acids 75<br />

and higher percentage gels (up to 20%), longer fixation and drying times are needed. For<br />

high percentage gels adding 10% glycerol to the fixative helps preserve the gel.<br />

23. If using 35-S, the use of enhancer screens is essential. Also, omit the cling film to ensure<br />

exposure. However, be careful to avoid moisture or the gel will stick to the film.<br />

References<br />

1. http://www.protocol-online.org<br />

2. Sanbrook, J., Fritsch, E. F., and Maniatis, T. (1989) Molecular Cloning: A Laboratory<br />

Manual. 2nd ed. Cold Spring Harbor Laboratory, Cold Spring, New York.<br />

3. <strong>Bartlett</strong>, J. M. S., Hulme, M. J., and Miller, W. R. (1996) Analysis of cAMP RI-alpha<br />

messenger-RNA expression in breast cancer—Evaluation of quantitative polymerase<br />

chain-reaction for routine use. Br. J. Cancer 73, 1538–1544.


76 <strong>Bartlett</strong>


Recovery of PCR Products from Gels 77<br />

18<br />

Technical Notes for the Recovery and Purification<br />

of PCR Products from Acrylamide Gels<br />

David Stirling<br />

1. Introduction<br />

Although the best way of obtaining pure polymerase chain reaction (PCR) product<br />

will always be to optimize reaction conditions to yield only one product, there are<br />

still circumstances where DNA has to purified from gels. Several good commercial<br />

products exist for the recovery of DNA from agarose. Here, we present a reliable<br />

method of recovering DNA from polyacrylamide gels.<br />

2. Materials<br />

1. Sterile scalpel blade.<br />

2. Microfuge tubes (0.5 and 1.5µL).<br />

3. Sterile narrow-gauge needle (23-guage).<br />

4. Tris-HCl EDTA buffer: 10 mM Tris-HCl, 10 mM EDTA, pH 8.0.<br />

5. 3 M sodium acetate (pH 4.0).<br />

6. 100% ethanol.<br />

7. 70% ethanol.<br />

3. Method<br />

1. Stain the gel with ethidium bromide, visualize the band of interest, and dissect with<br />

scalpel blade.<br />

2. Pierce hole in the bottom of a 0.5 ml Microfuge tube with narrow gauge needle.<br />

3. Place 0.5 mL of TE in a 1.5-mL microfuge tube.<br />

4. Place the acrylamide slice with the dissected DNA band into a 0.5-mL tube (with hole) and<br />

place inside the 1.5-mL tube (the lip will prevent it touching the bottom).<br />

5. Microfuge the two-tube combination at 12,000g for 5 min. The acrylamide will be forced<br />

through the hole in the bottom of the 0.5-mL tube into the TE.<br />

6. Discard the empty 0.5-mL tube.<br />

7. Cap the 1.5-mL tube, vortex for 30 s, then incubate for 2 h overnight at 37°C.<br />

8. Microfuge at 12,000g for 5 min.<br />

9. Transfer supernatant to fresh tube.<br />

10. Add 1/10 volume 3 M sodium acetate and 2 volumes 100% ethanol.<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

77


78 Stirling<br />

11. Store at –20°C for at least 30 min.<br />

12. Microfuge at 12,000g for 20 min at room temperature.<br />

13. Discard supernatant and wash pellet in 70% ethanol.<br />

14. Dry pellet and redissolve in 25 µL of TE.<br />

15. If concentration of DNA is insufficient for further processing, re-amplify 1 µL in a second<br />

round of PCR.


PCR Primer Design 81<br />

19<br />

PCR Primer Design<br />

David L. Hyndman and Masato Mitsuhashi<br />

1. Introduction<br />

The selection of primers for a given polymerase chain reaction (PCR) can determine<br />

the efficiency and specificity of the PCR. Although in many cases successful PCR<br />

primers have been selected with little understanding of the principles involved, PCR<br />

can often only be achieved by using primers that are designed appropriately. Here,<br />

we give general recommendations for PCR primer selection and various aspects to be<br />

considered when designing primers.<br />

2. General Primer Considerations<br />

2.1. Location<br />

The location of PCR primers is sometimes dictated by the purpose of the experiment.<br />

If the experiment is simply intended to identify the presence or absence of the sequence,<br />

then the location is of no consequence, so long as the amplification works well. If,<br />

however, the experiment is part of an assay for a particular allele of a gene, then the<br />

amplicon would be required to contain that region of interest.<br />

2.2. Amplicon Size<br />

In general, amplicons are from 100 to 1000 bp in length. The lower limit is caused<br />

by the typical need to be able to visualize the amplicon on an agarose gel. The upper<br />

limit of 1000 bp is the result of difficulties in amplifying large sequences. If an assay<br />

that does not require a minimum amplicon size is used, there is no theoretical minimum<br />

amplicon size.<br />

2.3. Guanine/Cytosine (G/C) Content<br />

Defined as the proportion of bases in the primer that are either G (guanine) or<br />

C (cytosine), good PCR primers are generally selected to have a G/C content between<br />

40 and 60%. However, there is no well-defined reason for this, only that it has been<br />

considered preferable.<br />

3. Considerations for Optimal PCR<br />

The issues concerning PCR primer design can be divided into two categories:<br />

efficiency and specificity. Both of these are important to consider in most applications,<br />

but often the factors that promote one of these will adversely affect the other.<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

81


82 Hyndman and Mitsuhashi<br />

3.1. Efficiency<br />

Efficiency can be viewed as the proportion of templates that are used to synthesize<br />

new strands with each round of PCR, assuming the PCR primers are in a high<br />

abundance. A situation in which efficiency is the primary concern is the amplification<br />

of a purified template in which there is no chance of nonspecific PCR amplifications,<br />

so the main issue is that of primer binding to its target.<br />

3.1.1. Melting Temperature<br />

Melting temperature, or Tm, is defined for a given DNA duplex as the temperature<br />

at which half of the strands are hybridized and half of the strands are not hybridized.<br />

The original definition of T m implied that the two complementary strands were in equal<br />

proportions. In cases where the two strands are not in equal proportions, such as a<br />

primer hybridizing to its target in PCR, the definition of Tm must be altered. Because<br />

in a PCR primer concentrations will be orders of magnitude higher than template<br />

concentrations, the interpretation is that the Tm is the temperature at which a primer is<br />

hybridized to half of the template strands. The T m of a given primer/template combination<br />

depends on primer concentration, template concentration, and salt concentration.<br />

Generally, the template concentration is considered to be negligible compared with the<br />

primer concentration, so the formula for calculating Tm is simplified such that it does<br />

not contain a parameter for template concentration.<br />

T m can be expressed as follows:<br />

∆H<br />

T m = ————— –273.15 + 16.6 log [Na + ] (1,2)<br />

∆S + Rln(c)<br />

The primer concentration is c; the ∆H and ∆S refer to the total enthalpy and entropy<br />

of hybridization, respectively; and [Na + ] is the sodium ion concentration but can<br />

refer to the total concentration of most monovalent cations (such as K + ). If there is<br />

Mg 2+ or other divalent cations such as Mn 2+ , the conversion is generally accepted<br />

as follows:<br />

Na + = 4 × [Mg 2+ ] 2 (3)<br />

3.1.2. Calculating T m with the Nearest Neighbor Model<br />

The accurate calculation of T m from a given sequence requires determining the<br />

∆H and ∆S of the hybridization. The most successful method for this is through the<br />

use of the nearest neighbor model (4–7). With the nearest neighbor model, every pair<br />

of adjacent base pairs makes a specific contribution to the overall ∆H and ∆S of the<br />

duplex. The total ∆H and ∆S are calculated by adding all of the component values<br />

plus a value for initiation of duplex formation. A table of ∆H and ∆S nearest neighbor<br />

values is shown in Table 1 (5).<br />

3.1.3. Efficient PCR Primers and T m<br />

The most important issue for designing efficient PCR primers is that they must bind<br />

to the target site efficiently under the conditions of the PCR. This generally means not<br />

only that they bind at the annealing temperature but that if the annealing temperature


PCR Primer Design 83<br />

Table 1<br />

Nearest Neighbor Thermodynamic Values<br />

for DNA Base Pairs<br />

Base Pair ∆H ∆S ∆G<br />

aa/tt 1–8.4 –23.6 –1.21<br />

at/ta 1– 6.5 –18.8 – 0.73<br />

aa/at 1– 6.3 –18.5 – 0.61<br />

ca/gt 1–7.4 –19.3 –1.38<br />

gt/ca 1–8.6 –231. –1.43<br />

ct/ga 1– 6.1 –16.1 –1.16<br />

ga/ct 1–7.7 –20.3 –1.46<br />

gc/cg –11.1 –28.4 –2.28<br />

gg/cc 1– 6.7 –15.6 –1.77<br />

is so low that the thermostable DNA polymerase is not active, the primer must be<br />

bound at the temperature at which the polymerase becomes active in order to begin<br />

extension.<br />

3.1.4. Typical Three-Step PCR<br />

As an illustration, let us examine a typical PCR cycle, where there is a dissociation<br />

step of 30 s at 95°C, an annealing step of 1 min at 37°C, and an extension step of 3<br />

min at 72°C. If the T m of the primer, in the conditions of the reaction, is 65°C, then<br />

the following will happen. As the temperature decreases from 95°C to 37°C, at some<br />

point the primer will hybridize to the template. At this temperature, however, the<br />

thermostable DNA polymerase may be almost totally inactive. As the temperature rises<br />

towards 72°C, at some point the polymerase becomes active and will start to extend the<br />

primer along the template. As the temperature rises above 65°C, some of the primers<br />

will dissociate if they haven’t been extended. Those that have extended sufficiently,<br />

however will form a more stable duplex as a result of their added base pairs from the<br />

extension, and they will not dissociate before reaching 72°C. During the extension at<br />

72°C, the primers still hybridized will be fully extended to generate new strands.<br />

Let’s now imagine that with this scenario the T m of the primer is 75°C. In this<br />

case, as the temperature rises from annealing to extension, even primers that have<br />

not extended will not dissociate. In this scenario, there will be almost total extension<br />

of every possible target strand. Therefore, this PCR would have a high degree of<br />

efficiency.<br />

3.1.5. Two-Step PCR<br />

PCR is sometimes performed in two steps without a discrete annealing step. In this<br />

case, the annealing takes place at the same temperature as the extension. This requires<br />

that the primers will hybridize to some degree at the extension temperature. If the<br />

extension temperature is 72°C, then a primer with a T m of 72°C would be an efficient<br />

primer. Because the definition of T m is the point at which half of the templates are in<br />

a duplex, a short time after reaching 72°C, half of the templates will be bound by a<br />

primer. Shortly after that, a large percentage of those templates will be extended by


84 Hyndman and Mitsuhashi<br />

Fig. 1. Hairpin structures.<br />

the polymerase and thereby taken out of the equilibrium between bound and unbound<br />

templates. Because the primer concentration will essentially be unchanged, half of the<br />

remaining templates will then be bound by primers and extended. In this way, most<br />

templates can be extended if the extension is done at the T m of the primer.<br />

If, however, the T m of the primer is a few degrees below the extension temperature,<br />

only a small percentage of the primers will be hybridized, and the PCR will not be<br />

efficient. Therefore, very efficient and specific PCR can be performed with two-step<br />

cycling, but a sufficiently high primer T m is very important.<br />

3.1.6. Hairpins<br />

A hairpin is a structure formed by a single DNA molecule in which a portion on<br />

one part of the DNA hybridizes to a complementary portion within the same DNA<br />

strand, forming a structure resembling a hairpin (Fig. 1A). When a PCR primer forms<br />

a hairpin, it adversely affects the primer’s ability to bind and extend at the target site.<br />

In the worst case, the hairpin includes a base pair of the 3′-end and an overhang of the<br />

5′-end (Fig. 1B). Such a structure allows the extension by DNA polymerase along the<br />

primer and will result in the formation of a primer that will not be complementary to<br />

the template and will not be extended if hybridized (Fig. 1C). In addition to removing<br />

primers from the mixture, this also will prevent native primers from binding as target<br />

sites that are bound by the extended primers. To avoid this, primers should be selected<br />

that do not have any possible hairpin structures if possible.<br />

3.1.7. Primer-Dimer Formation<br />

The hybridization of two primers together is referred to as a primer-dimer (Fig. 2A).<br />

There are two possibilities for these, homodimers and heterodimers. Homodimers are<br />

formed from the hybridization of the same species of primer together. Heterodimers<br />

are the duplex of two different primer sequences hybridizing together. The result of<br />

either of these is that the primers will not be as efficient in hybridizing to the target.


PCR Primer Design 85<br />

Fig. 2. Dimer structures.<br />

As with hairpins, the worst case is that in which the 3′-end of one of the primers is<br />

base paired and there is a 5′ overhang (Fig. 2B). In this case, the primer will extend,<br />

using the other primer as a template, rendering the extended primer unable to prime<br />

the desired template (Fig. 2C). Even worse than with hairpins, this situation leads to<br />

amplification of the primer dimers and rapid depletion of useable primers. To prevent<br />

this, primer pairs should be chosen such that primer-dimer formation is minimal.<br />

3.1.8. 3′-Terminal Stability<br />

3′-terminal stability can loosely be defined as the relative hybridization strength<br />

of the 3′-end of the primer. If the 3′-end of the primer has a low stability, it may not<br />

efficiently prime because of the transient fraying of the end of the duplex. Therefore, a<br />

higher 3′-terminal stability will improve priming efficiency. As will be mentioned later,<br />

however, this high stability can have an undesirable affect on specificity.<br />

3.2. Specificity<br />

Specificity can generally be defined as the tendency for a primer to hybridize to its<br />

intended target and not to other, nonspecific, targets. There are a few ways in which<br />

poor specificity can impair PCR. First, if primers are hybridizing to many locations<br />

nonspecifically, they will not be available to prime the target sequence. Second, if such<br />

nonspecific hybridization were to occur, priming could also occur at those nonspecific<br />

locations, which would effectively remove the primers from the reaction permanently.<br />

Finally, by priming nonspecifically, it can be possible to generate aberrant amplicons.<br />

This will not only obfuscate an assay for successful PCR, but will very rapidly consume<br />

the primers to remove them from the reaction for amplifying the intended target.<br />

3.2.1. Specificity, T m , and PCR Conditions<br />

With respect to the annealing and extension temperatures chosen for the PCR<br />

reaction, there is a balance that must be reached between considerations of efficiency<br />

and specificity. As discussed previously, a more efficient PCR will result from having<br />

primer T m equal to or above the extension temperature of the reaction. However,<br />

having a primer with a high T m can often result in poor specificity. In such<br />

cases, partial hybridization of the primer may be likely and extension can occur from<br />

nonspecific sites. Such issues are less critical if highly specific primers can be selected<br />

as discussed here.


86 Hyndman and Mitsuhashi<br />

Fig. 3. Hybridization simulation data.<br />

3.2.2. Hybridization Simulation<br />

The most precise way to view the specificity of a PCR primer is by hybridization<br />

simulation (8). Hybridization simulation is the computer simulation of a hybridization<br />

of a primer with a specified database. This in silico analysis will identify all hybridization<br />

sites within the database for a candidate primer, allowing the user to select primers<br />

that will be the most specific.<br />

It is important to realize that hybridization simulation is qualitatively different<br />

from a homology or similarity analysis (9). Hybridization simulation uses a thermodynamic<br />

model with nearest neighbor values to calculate the mismatch Tm of hybridization<br />

for all hybridization sites. An example of hybridization simulation data is<br />

shown in Fig. 3.<br />

Currently, hybridization simulation is only available from a single commercially<br />

available program, the HYBsimulator . HYBsimulator allows for the screening of a large<br />

set of candidate primers and selection based on the hybridization simulation data.<br />

3.2.3. Statistical Determination of Specificity<br />

Various mathematical models exist in which the specificity of a given primer can<br />

be estimated based on the frequency of its constituent smaller sequences. One such<br />

method uses a table of frequencies of 6 mers found within a given genomic database


PCR Primer Design 87<br />

(10,11). The entire statistical frequency of the entire oligonucleotide is calculated<br />

based on the constituent 6-mers by starting with the 5′ terminal 6-mer frequency and<br />

multiplying the relative frequency of the 5-mer on the 3′-end of that 6 mer having the<br />

next nucleotide. This is repeated until the end of the oligonucleotide is reached.<br />

For example, to calculate the frequency (f) of the 8-mer: CATAGCCT<br />

f(CATAGCCT) =<br />

4 f(ATAGCC)<br />

f(CATAGC) × ———————————————————————<br />

f(ATAGCT) + f(ATAGCG) + f(ATAGCC) + f(ATAGCT)<br />

f(CATAGC)<br />

4 f(TAGCCT)<br />

f(CATAGC) × ———————————————————————<br />

f(TAGCCA) + f(TAGCCG) + f(TAGCCC) + f(TAGCCT)<br />

where f(CATAGC) denotes the frequency of the 6-mer, CATAGC.<br />

3.2.4. 3′-Terminal Effects<br />

Partial hybridization of the primer at the 3′-terminus can permit extension by DNA<br />

polymerase. This could result in depletion of primers as well as possible nonspecific<br />

amplification; therefore, this type of partial hybridization should be minimized as<br />

much as possible.<br />

There are two considerations for decreasing the chance of partial hybridization of<br />

the 3′-terminal: frequency and stability.<br />

3.2.5. 3′-Terminal Frequency<br />

If the 3′-terminal region has a sequence that has many occurrences in the DNA<br />

that will be in the reaction, then the likelihood of partial hybridization is greater. To<br />

minimize this, the primers can be selected such that the 3′-terminal region does not<br />

have a high frequency of occurrence within the genome of interest.<br />

3.2.6. 3′-Terminal Stability<br />

If the 3′-terminal region has a strong hybridization energy, then 3′-terminal partial<br />

hybridizations will be relatively more stable. More stable 3′-terminal hybridization will<br />

allow more false priming. Therefore if the primers are chosen such that the 3′-terminal<br />

has a low hybridization strength (also referred to as terminal stability), the primer is<br />

less likely to be priming as a result of such partial hybridization.<br />

Note that in the efficiency section above, a stronger 3′-terminal stability is said<br />

to improve efficiency. Whether one should select primers with high or low terminal<br />

stability would depend on factors, such as the nature of the experiment (i.e., whether<br />

there will be a large amount of other DNA) and the nature of the gene (whether highly<br />

specific primers can be found).<br />

3.2.7. Specificity Within the Target Sequence<br />

If the primer is able to partially hybridize to a nonspecific region of the template,<br />

particularly undesirable effects can occur. If the nonspecific hybridization allows<br />

extension along the template such that a PCR product can be formed in conjunction<br />

with one of the primers binding at one of the actual binding sites, nonspecific amplicons


88 Hyndman and Mitsuhashi<br />

can be generated. This problem is compounded if the nonspecific binding site is within<br />

the amplicon itself. For this reason, primers should be checked for nonspecific alternate<br />

hybridization sites within the target sequence.<br />

4. Selecting Primers for Multiplex PCR<br />

Multiplex PCR, in which several primer sets amplify several amplicons in the same<br />

reaction, add a degree of complexity to designing optimal primers. The additional<br />

issues to consider are those of possible heterodimer formation between all of the<br />

candidate primers and possible alternate hybridization sites within any of the target<br />

sequences. Some of the available primer design software provides functions for these<br />

types of designs.<br />

5. Primer Design Software<br />

Several programs are available for PCR primer design. As mentioned above, we<br />

think HYBsimulator is the most powerful such program and does provide PCR primer<br />

selection based on all criteria mentioned in this chapter. Other popular programs<br />

are Oligo and Primer Premier , which provide a subset of these functions but are<br />

slightly easier to use.<br />

References<br />

1. Breslauer, K. J., Frank, R., Blocker, H., and Marky, L. (1986) Predicting DNA duplex<br />

stability from the base sequence. Proc. Natl. Acad. Sci. USA 83, 3746–3750.<br />

2. Freir, S. M., Kierzed, R., Jaeger, J. A., Sugimoto, N., Caruthers, M. H., Neilson, T., and<br />

Turner, D. H. (1986) Improved free-energy parameters for predictions of RNA duplex.<br />

<strong>Bio</strong>chemistry 83, 9373–9377.<br />

3. Wetmer, J. (1991) DNA probes: Applications of the principles of nucleic acid hybridization.<br />

Crit. Rev. <strong>Bio</strong>chem. Mol. <strong>Bio</strong>l. 26, 227–259.<br />

4. Sugimoto, N., Nakano, S., Katoh, M., Matsumura, A., Nakamuta, H., Ohmichi, T., et al.<br />

(1995) Thermodynamic parameters to predict stability of RNA/DNA hybrid duplexes.<br />

<strong>Bio</strong>chemistry 34, 11,211–11,216.<br />

5. SantaLucia, J., Allawi, H. T., and Seneviratne, P. A. (1996) Improved nearest-neighbor<br />

parameters for predicting DNA duplex stability. <strong>Bio</strong>chemistry 35, 3555–3562.<br />

6. SantaLucia, J., Kierzed, R., and Turner, D. H. (1990) Effects of GA mismatches on the<br />

structure and thermodynamics of RNA internal loops. <strong>Bio</strong>chemistry 29, 8813–8819.<br />

7. Sugimoto, N., Kierzed, R., Freier, S. M., and Turner, D. H. (1986) Energetics of internal GC<br />

mismatches in ribooligonulceotide helix. <strong>Bio</strong>chemistry 25, 5755–5759.<br />

8. Hyndman, D., Cooper, A., Pruzinsky, S., Coad, D., and Mitsuhashi, M. (1996) Software<br />

to determine optimal oligonucleotide sequences based on hybridization simulation data.<br />

<strong>Bio</strong>Techniques 20, 1090–1096.<br />

9. Altschul, S. F., Gish, W., Miller, W., Myers, E. W., and Lipman, D. J. (1990) Basic local<br />

alignment search tool. J. Mol. <strong>Bio</strong>l. 215, 403– 410.<br />

10. Han, J., Hsu, C., Zhu, Z., Longshore, J., and Finley, H. (1994) Over-representation of<br />

the disease associated (CAG) and (CGG) repeats in the human genome. Nucleic Acids<br />

Res. 22, 1735–1740.<br />

11. Han, J., Zhu, Z., Hsu, C., and Finley, W. (1994) Selection of antisense oligonucleotides on<br />

the basis of genomic frequency of the target sequence. Antisense Res. Devel. 4, 53–65.


Optimization of PCRs 89<br />

20<br />

Optimization of Polymerase Chain Reactions<br />

Haiying Grunenwald<br />

1. Introduction<br />

The polymerase chain reaction (PCR) is a powerful method for fast in vitro enzymatic<br />

amplifications of specific DNA sequences. PCR amplifications can be grouped into<br />

three different categories: standard PCR, long PCR, and multiplex PCR. Standard PCR<br />

involves amplification of a single DNA sequence that is less than 5 kb in length and<br />

is useful for a variety of applications, such as cycle sequencing, cloning, mutation<br />

detection, etc. Long PCR is used for the amplification of a single sequence that is longer<br />

than 5 kb and up to 40 kb in length. Its applications include long-range sequencing;<br />

amplification of complete genes; PCR-based detection and diagnosis of medically<br />

important large-gene insertions or deletions; molecular cloning; and assembly and<br />

production of larger recombinant constructions for PCR-based mutagenesis (1,2). The<br />

third category, multiplex PCR, is used for the amplification of multiple sequences<br />

that are less than 5 kb in length. Its applications include forensic studies; pathogen<br />

identification; linkage analysis; template quantitation; genetic disease diagnosis; and<br />

population genetics (3–5). Unfortunately, there is no single set of conditions that is<br />

optimal for all PCR. Therefore, each PCR is likely to require specific optimization for<br />

the template/primer pairs chosen. Lack of optimization often results in problems,<br />

such as no detectable PCR product or low efficiency amplification of the chosen<br />

template; the presence of nonspecific bands or smeary background; the formation of<br />

“primer-dimers” that compete with the chosen template/primer set for amplification;<br />

or mutations caused by errors in nucleotide incorporation. It is particularly important<br />

to optimize PCR that will be used for repetitive diagnostic or analytical procedures<br />

where optimal amplification is required. The objective of this chapter is to discuss the<br />

parameters that may affect the specificity, fidelity, and efficiency of PCR, as well as<br />

approaches that can be taken to achieve optimal PCR amplifications.<br />

Optimization of a particular PCR can be time consuming and complicated because<br />

of the various parameters that are involved. These parameters include the following: (1)<br />

quality and concentration of DNA template; (2) design and concentration of primers;<br />

(3) concentration of magnesium ions; (4) concentration of the four deoxynucleotides<br />

(dNTPs); (5) PCR buffer systems; (6) selection and concentration of DNA polymerase;<br />

(7) PCR thermal cycling conditions; (8) addition and concentrations of PCR<br />

additives/cosolvents; and (9) use of the “hot start” technique. Optimization of PCR<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

89


90 Grunenwald<br />

may be affected by each of these parameters individually, as well as the combined<br />

interdependent effects of any of these parameters.<br />

2. Materials<br />

1. Template DNA (e.g., plasmid DNA, genomic DNA).<br />

2. Forward and reverse PCR primers.<br />

3. MgCl 2 (25 mM).<br />

4. dNTPs (a mixture of 2.5 mM dATP, dCTP, dGTP, and dTTP).<br />

5. 10× PCR buffer: 500 mM KCl, 100 mM Tris-HCl, pH 8.3, 25°C.<br />

6. Thermal stable DNA polymerase (e.g., Taq DNA polymerase).<br />

7. PCR additives/cosolvents (optional; e.g., betaine, glycerol, DMSO, formamide, bovine<br />

serum albumin, ammonium sulfate, polyethylene glycol, gelatin, Tween-20, Triton X-100,<br />

β-mercaptoethanol, or tetramethylammonium chloride).<br />

3. Methods<br />

3.1. Setting Up PCR<br />

The common volume of a PCR is 10, 25, 50, or 100 µL. Although larger volumes are<br />

easier to pipet, they also use up a larger amount of reagents, which is less economical.<br />

All of the reaction components can be mixed in together in a 0.5-mL PCR tube in<br />

any sequence except for the DNA polymerase, which should be added last. It is<br />

recommended to mix all the components right before PCR cycling. Although it is not<br />

necessary to set up the PCR on ice, some published protocols recommend it.<br />

For each PCR, the following components are mixed together:<br />

1. Template DNA (1–500 ng).<br />

2. Primers (0.05–1.0 µM).<br />

3. Mg 2+ (0.5–5 mM).<br />

4. dNTP (20–200 µM each).<br />

5. 1× PCR buffer: 1 mM Tris-HCl and 5 mM KCl.<br />

6. DNA polymerase (0.5–2.5 U for each 50 µL of PCR).<br />

As a real-life example, the following PCR was set up to amplify the cII gene from<br />

bacteriophage lambda DNA (total volume = 50 µL):<br />

1. 1 µL of 1 ng/µL lambda DNA (final amount = 1 ng).<br />

2. 1 µL of 50 µM forward PCR primer (final concentration = 1 µM).<br />

3. 1 µL of 50 µM reverse PCR primer (final concentration = 1 µM).<br />

4. 5 µL of 25 mM MgCl 2 (final concentration = 2.5 mM).<br />

5. 4 µL of 2.5 mM dNTPs (final concentration = 200 µM).<br />

6. 5 µL of 10× PCR buffer (final concentration = 1×).<br />

7. 0.25 µL of 5 U/µL Taq DNA polymerase (final amount = 1.25 U).<br />

3.2. PCR Cycling<br />

A common PCR cycling program usually starts with an initial dissociation step at<br />

92 to 95°C for 2 to 5 min to ensure the complete separation of the DNA strands. Most<br />

PCR will reach sufficient amplification after 20 to 40 cycles of strand denaturation<br />

at 90 to 98°C for 10 s to 1 min, primer annealing at 55 to 70°C for 30 s to 1 min,<br />

and primer extension at 72 to 74°C for 1 min per kilobase of expected PCR product.<br />

It is suggested that a final extension step of 5 to 10 min at 72°C will ensure that


Optimization of PCRs 91<br />

all amplicons are fully extended, although no solid evidence proves that this step is<br />

necessary. For example, the cycling program used to amplify the previously described<br />

lambda cII gene is as follows: initial denaturation for 4 min at 94°C, followed by 30<br />

cycles of 1 min at 94°C, 1 min at 55°C, and 1 min at 72°C, then held at 4°C.<br />

3.3. Verifying PCR Amplification<br />

To measure the success of a PCR amplification, 5 to 10 µL of the final PCR product<br />

is run on a 1 or 2% agarose gel and visualized by staining with ethidium bromide. The<br />

critical questions are as follows: (1) Is there a band on the gel? (2) Is the band at the<br />

expected size? (3) Are there any nonspecific bands beside the expected PCR band on<br />

the gel? (4) Is there smear on the gel? A successful PCR amplification should display a<br />

single band with the expected size without nonspecific bands and smear.<br />

4. Notes<br />

1. The quality and concentration of DNA templates can directly affect the outcome of PCR<br />

amplifications. To achieve satisfactory amplification, certain baseline conditions may be<br />

used as a starting point for optimizing a PCR amplification. For a typical PCR, 10 4 to 10 7<br />

molecules of template DNA is recommended. For long PCR (>5 kb), 10 7 to 10 8 molecules<br />

of a high copy number template DNA (e.g., 1–10 ng of lambda DNA) is recommended.<br />

For amplification from genomic DNA, use 100 to 500 ng of template DNA. In multiplex<br />

PCR, two- to fivefold more DNA template than what is needed for a typical PCR should<br />

be used.<br />

There are several methods for purifying DNA for PCR amplification, including commercially<br />

available kits, as well as standard methods (6). Long PCR amplification has<br />

the most stringent requirement for the quality of template DNA. Care should be taken to<br />

prevent template DNA damages from nicking, shearing, and depurination (7). Analysis by<br />

pulsed-field agarose gel electrophoresis is typically recommended for template DNA used<br />

in long PCR to assure its purity and integrity.<br />

2. Appropriate primer design, as well as use of the proper primer concentration, is critical<br />

for successful PCR amplification. There are a variety of computer programs available for<br />

designing primers and they vary significantly in selection criteria, comprehensiveness, and<br />

user-friendliness (8–11). The purpose of primer design is to achieve a balance between<br />

the specificity and efficiency of an amplification. Specificity defines how frequently<br />

mispriming occurs, whereas efficiency represents the increase of the amount of PCR<br />

product over a given number of cycles. The following guidelines should be considered<br />

when designing primers. The optimal primer size is usually between 18 and 28 bases.<br />

Shorter primers are generally less specific but may result in more efficient PCR, whereas<br />

longer primers improve specificity yet can be less efficient. Primers from both directions<br />

should have melting temperatures (T m , defined as the dissociation temperature of the<br />

primer/template duplex) that are within 2 to 5°C of each other. This will assure that the<br />

proper annealing temperature for both primers is achieved. For primers shorter than 20<br />

bases, an estimate of T m can be calculated as T m = 4 (G + C) + 2 (A + T) (12). However,<br />

for longer primers, correct estimation of T m requires a “nearest-neighbor” calculation,<br />

which takes into account thermodynamic parameters of a chosen primer and is used by<br />

most of the available computer programs for primer design (13,14). Avoid complementary<br />

sequences within a primer or between the two primers. This will reduce formation of<br />

primer-dimers that can compete with the amplification of the desired PCR product, as well<br />

as the formation of secondary structures within a primer. Primers with T m higher than


92 Grunenwald<br />

50°C will generally provide specific and efficient amplifications. For long PCR, a T m of 62<br />

to 70°C is recommended. In addition, a GC content of 40 to 60% is desirable for primers<br />

because this assures a higher T m and therefore increases specificity. If AT content is high,<br />

use primers of 28 to 35 bases in length for long PCR. Also, avoid continuous stretches<br />

of purines or pyrimidine, as well as multiple repeats of thymidine residues at the 3′ end<br />

of the primer. It is extremely critical when designing primers for multiplex PCR to make<br />

sure that all primers have similar melting temperatures and do not contain sequences<br />

complementary to each other. Concentrations of primers used in PCR will also influence<br />

amplification specificity and efficiency. In general, primer concentrations between<br />

0.05 to 1.0 µM (or 0.5–100 pmol) are routinely used in 100 µL of PCR depending on the<br />

specific application. Higher primer concentrations can result in nonspecific priming and<br />

formation of primer-dimers, whereas lower primer concentrations may adversely affect<br />

PCR efficiency. Use the primers at a 11 concentration ratio to assure specificity and<br />

efficiency of amplifications. Lower concentrations of primers are more desirable for<br />

multiplex PCR because of the number of primer sets present in the reaction.<br />

3. Magnesium concentration is critical to the success of PCR amplification because it may<br />

affect DNA polymerase activity and fidelity, DNA strand denaturation temperatures of both<br />

template and PCR product, primer annealing, PCR specificity, and primer-dimer formation.<br />

Excess magnesium results in accumulation of nonspecific amplification products seen<br />

as multiple bands on an agarose gel, whereas insufficient magnesium results in reduced<br />

yield of the desired PCR product. Common magnesium concentrations used in PCR are<br />

between 0.5 to 5 mM. It is important to optimize the magnesium concentration used for<br />

each individual PCR because DNA polymerases require free magnesium for their activity<br />

in addition to that bound by template DNA, primers, and dNTPs. Also, trace amounts<br />

of EDTA or other chelators may be present in primer stock solutions or template DNA.<br />

Therefore, each PCR should contain 0.5 to 2.5 mM magnesium over the total dNTP<br />

concentrations (15). A simple way to optimize magnesium concentration is to first perform<br />

a series of reactions in which the magnesium concentration is varied between 0.5 and 5 mM<br />

in 0.5-mM increments. After the concentration range is narrowed, perform a second round<br />

of reactions and vary the magnesium concentration in 0.2- to 0.3-mM increments.<br />

4. Concentration of dNTPs can affect the yield, specificity, and fidelity of a PCR amplification.<br />

Concentrations of 20 to 200 µM of each dNTP has been used to obtain successful<br />

PCR amplifications. Stock solutions of each dNTP are adjusted to pH 7.0 and diluted<br />

to a 10 mM final concentration. Commercially available premixed dNTP solutions with<br />

concentrations of 2.5 mM or individual dNTP stock solutions of 10 mM may be used.<br />

Lower concentrations of dNTPs minimize mispriming and reduce the likelihood of<br />

extending misincorporated nucleotides, which in turn increase specificity and fidelity<br />

of PCR amplifications (16). Because dNTPs are typically added in excess to a PCR,<br />

one should determine the lowest dNTPs concentration appropriate for the length and<br />

composition of the target sequence. Although 250 µM of each of the dNTPs appears to be<br />

sufficient for long PCR (17), amplifications of sequences longer than 20 kb may require<br />

dNTP concentrations as high as 400 to 500 µM each in a given 50-µL reaction. It is critical<br />

not to use a large excess of dNTP because higher dNTP concentrations increase the error<br />

rate of DNA polymerases. In fact, millimolar concentrations of dNTPs actually inhibit<br />

Taq DNA polymerase (18).<br />

5. Standard PCR amplifications using Taq DNA polymerase are performed in 10 mM Tris-<br />

HCl (pH 8.3–8.4 at 20–25°C) and 50 mM KCl. For Tth and Tfl DNA polymerases, and<br />

DNA polymerases with proofreading activity (for example, Pwo, Pfu, Tli, and Vent DNA<br />

polymerases (New England <strong>Bio</strong>labs), a buffer system of 50 mM Tris-HCl (pH 9.0 at 25°C)<br />

and 20 mM (NH 4 ) 2 SO 4 is normally used. These standard buffer systems are available


Optimization of PCRs 93<br />

commercially and have been shown to produce satisfactory PCR amplifications in most<br />

cases. However, long PCR requires a different buffer system. For example, 20 to 25 mM<br />

Tricine (pH 8.7 at 25°C) and 80 to 85 mM potassium acetate (pH 8.3–8.7 at 25°C)<br />

(19), as well as 25 mM Tris-HCl (pH 8.9 at 25°C) and 100 mM KCl (2), have been<br />

used successfully in long PCR amplifications when used in conjunction with rTth DNA<br />

polymerase (Perkin–Elmer). The use of less temperature-sensitive buffers, such as Tricine,<br />

may enhance the ability to obtain long PCR amplifications.<br />

6. The most common enzyme used for PCR amplification is Taq DNA polymerase because<br />

of its thermostability and processivity (i.e., the number of nucleotides replicated before<br />

the enzyme dissociates from the DNA template). It was originally purified from the<br />

gram-negative thermophilic bacterium Thermus aquaticus (20). Highly purified Taq DNA<br />

polymerase exhibits a temperature optimum of 75 to 80°C (21). The half-life of Taq<br />

DNA polymerase is 40 min at 95°C, which is sufficient to remain active over 30 or more<br />

cycles, during which the enzyme is transiently exposed to extremely high denaturation<br />

temperatures. Taq DNA polymerase has an extension rate of 35 to 100 nucleotides per<br />

second at 72°C (16), which is the most common extension temperature for PCR amplifications.<br />

A recommended concentration range for Taq DNA polymerase is between 1 and<br />

2.5 units per 100 µL of PCR. However, different concentrations of Taq DNA polymerase<br />

may be required with respect to individual target template sequences or primers. Increasing<br />

the amount of Taq DNA polymerase beyond the 2.5 units/reaction can in some cases<br />

increase PCR efficiency. However, adding more Taq DNA polymerase can sometimes<br />

increase the yield of nonspecific PCR products at the expense of the desired product. When<br />

optimizing the Taq DNA polymerase concentration for a particular PCR, testing a range<br />

of 0.5 to 5 units per 100 µL of reaction in 0.5-unit increments is recommended, followed<br />

by analysis of the PCR products by gel electrophoresis to determine the amplification<br />

specificity and efficiency.<br />

The other important property of Taq DNA polymerase is its fidelity, which is measured<br />

as error rate. The error rate for Taq DNA polymerase, which lacks proofreading 3′ → 5′<br />

exonuclease activity, is estimated at approx 1 to 2 × 10 –5 errors (or mutation frequency)<br />

per nucleotide per duplication (22–24). For many applications, this does not present<br />

any problems. However, for some sequencing, cloning, and long PCR applications, it is<br />

essential to have few, or no incorporation errors. In situations where “high fidelity” is<br />

required, DNA polymerases with 3′ → 5′ proofreading activity (for example, Pfu or Pwo<br />

DNA polymerases) are recommended. The estimated error rates for these proofreading<br />

enzymes is approx 1 to 2 × 10 –6 errors per nucleotide per duplication (23,24), representing a<br />

10-fold improvement over standard Taq DNA polymerase. It is important to note that these<br />

proofreading enzymes with lower error rates also have lower extension rates, resulting<br />

in lower PCR efficiency. Therefore, more amplification cycles are required to achieve<br />

adequate amount of amplified DNA.<br />

DNA polymerase characteristics, such as extension rate, processivity, fidelity, thermostability,<br />

and thermal activity profile, are important in long PCR. PerkinElmer ’s<br />

rTth DNA polymerase (the recombinant form of the DNA polymerase from T. thermophilus)<br />

has been shown to perform consistent long PCR at 0.5 to 2.5 units per 50 µL<br />

of reaction (17). Use of a mixture of Taq DNA polymerase with a proofreading enzyme,<br />

such as Pfu DNA polymerase, at a 201 ratio (2.5 units per 50 µL of reaction) will also<br />

enhance the reliability of long PCR amplifications.<br />

7. Optimization of PCR thermal cycling conditions includes determination of cycle number,<br />

the temperature and incubation time period for template denaturation, primer annealing,<br />

and primer extension. The optimum number of cycles depends mainly on the starting<br />

concentration of template DNA. Because many PCRs start with very limiting amount


94 Grunenwald<br />

of template DNA, a sufficient number of cycles are required to achieve satisfactory<br />

amplifications. However, PCR amplification is not an unlimited process. A common<br />

mistake is to execute too many cycles. The exponential amplification of PCR will continue<br />

up until the point when the product reaches about 10 –8 M (about 10 12 molecule in a 100-µL<br />

reaction). The reaction enters a linear phase where exponential accumulation of the product<br />

is attenuated. This is termed the plateau effect (25). In most PCRs, amplifications plateau<br />

after about 20 to 40 cycles. The other major limitation of standard PCR is the amount of<br />

DNA polymerase included in the reaction. The combination of thermal inactivation of the<br />

DNA polymerase after each denaturation step, reduction in denaturation efficiency, and<br />

the reduced efficiency of primer annealing (caused by increasing competition from the<br />

template), will cause the reaction to terminate. Although too few cycles of PCR result<br />

in low product yield, most PCR amplifications are performed for no more than 20 to<br />

40 cycles.<br />

It is often helpful to precede the first cycling denaturation step with an initial dissociation<br />

step at 92 to 95°C for 2 to 5 min to ensure the complete separation of the DNA strands.<br />

Template denaturation temperatures range 90 to 98°C. The duration of denaturation ranges<br />

from 10 s to 1 min. Although it only takes a few seconds to denature DNA at its strandseparation<br />

temperature, it is appropriate to use a higher denaturation temperature and a<br />

longer incubation time for some templates, such as those templates with high GC content,<br />

to achieve complete denaturation. Although a higher temperature and a longer incubation<br />

period result in a more complete denaturation of the DNA template, it can also cause<br />

depurination of the DNA template, which in turn reduces amplification efficiency. It is<br />

also important to note that higher denaturation temperatures will reduce the amount of<br />

active DNA polymerase available for amplification. The half-life of Taq DNA polymerase<br />

activity is more than 2 h at 92.5°C, 40 min at 95°C, and 5 min at 97.5°C.<br />

The optimal primer annealing temperature for a particular PCR amplification depends<br />

on the base composition, nucleotide sequence, length, and concentration of the primers<br />

(26). A typical primer annealing temperature is 5°C below the calculated Tm of the primers.<br />

Annealing temperatures from 55 to 70°C generally yield the best results. Increasing the<br />

annealing temperature enhances discrimination against incorrectly annealed primers and<br />

reduces mis-extension of incorrect nucleotides at the 3′ end of primers. Therefore, a<br />

higher annealing temperature increases amplification specificity. For these reasons, when<br />

using primers with higher T m s such as for long PCR, a higher annealing temperature (i.e.,<br />

60–70°C) should be used. It is also important to note that at typical primer concentrations<br />

of 200 µM each, annealing requires only a few seconds. However, incubation times from<br />

30 s to 1 min are generally recommended to assure successful primer annealing.<br />

Primer extension time depends on the length and concentration of the target sequence,<br />

as well as the extension temperature. Taq DNA polymerase extends at a rate of 0.25 nucleotides<br />

per second at 22°C, 1.5 nucleotides per second at 37°C, 24 nucleotides per second<br />

at 55°C, greater than 60 nucleotides per second at 70°C, and 150 nucleotides per second<br />

at 75 to 80°C (18). Therefore, at the commonly chosen extension temperature of 72°C,<br />

Taq DNA polymerase is expected to extend at the rate of greater than 3500 nucleotides<br />

per minute. Thus, as a general rule, an extension time of 1 min per kilobase is more<br />

than sufficient to generate the expected PCR product. For PCR products up to 2 kb in<br />

length, an extension time of 1 min at 72°C is sufficient. A final extension step of 5 to<br />

10 min at 72°C may be added in order to ensure that all amplicons are fully extended.<br />

In addition to the conventional three-step cycling programs, two-step cycling programs<br />

that combine primer annealing and extension in one step are also widely used. Annealing<br />

and extension can be combined because most thermostable DNA polymerases can actively<br />

extend off the primers over the entire range of commonly chosen annealing and extension


Optimization of PCRs 95<br />

temperatures. Two-step cycling programs are generally applied when a high annealing<br />

temperature is used, such as 65 to 70°C. Because a higher annealing temperature improves<br />

amplification specificity, it is argued by some investigators that better PCR results may be<br />

obtained using a two-step cycling program (27). Here is an example of a typical two-step<br />

cycling program: initial denaturation at 94°C for 2 min, followed by 30 cycles of 1 min at<br />

94°C and 1 min at 68°C, then held at 4°C.<br />

It is reasonable to follow the above rationale in defining optimal cycling conditions<br />

for long PCR. However, there are a few rules that must be considered when performing<br />

long PCR. Use moderate denaturation temperatures and short incubation time periods<br />

to maintain the integrity of the long DNA templates. Choose primers with a high Tm<br />

so that a higher annealing temperature can be used to increase specificity. Two-step<br />

cycling programs are more frequently used than three-step cycling programs. For example,<br />

denaturation at 92 to 95°C for 10 to 30 s, followed by annealing and extension at 65 to<br />

68°C for 1 min per kilobase will increase the probability of obtaining the desired product.<br />

Programming an increase in extension time automatically in later cycles may also improve<br />

the yields of the amplification. Because DNA polymerases extend primers discontinuously<br />

through a succession of reactions, increasing the extension time in each of the later PCR<br />

cycles could increase the likelihood of synthesizing long PCR products. For example,<br />

perform the extension at 1 min per kilobase for the first 10 cycles, then lengthen the<br />

extension time 10 to 20 s for each of the next 20 cycles. Here is an example of a typical<br />

cycling program used to amplify a 10-kb PCR product: initial denaturation at 94°C for<br />

2 min, followed by 16 cycles of 30 s at 94°C and 10 min at 68°C, then 12 cycles of 30 s at<br />

94°C and 10 mins at 68°C with a 15-s extension per cycle, then held at 4°C.<br />

8. The majority of PCR amplifications can be successfully performed after optimizing the<br />

above parameters. However, there are some PCR amplifications, for example, those using<br />

templates with high guanine/cytosine content or stable secondary structures, that may still<br />

amplify inefficiently, resulting in little or no desired product and/or nonspecific products.<br />

Incorporation of the nucleotide analog 7-deaza-2′-deoxyguanosine triphosphate (c7dGTP)<br />

in addition to deoxyguanosine triphosphate (dGTP) helps destabilize secondary structures<br />

of DNA and reduces the formation of nonspecific products (28). However, the most<br />

effective and frequently used strategy is addition of various organic additives or cosolvents.<br />

The most commonly used cosolvents and their concentration ranges are: dimethyl<br />

sulfoxide (DMSO; 1–10%), glycerol (5–20%), formamide (1.25–10%), bovine serum<br />

albumin (10–100 µg/mL), ammonium sulfate (NH 4 ) 2 SO 4 ; 15–30 mM), polyethylene<br />

glycol (5–15%), gelatin (0.01%), non-ionic detergents (such as Tween 20 and Triton<br />

X-100; 0.05–0.1%), β-mercaptoethanol, tetramethylammonium chloride(TMAC), and<br />

N,N,N-trimethyglycine (betaine) (1–3 M) (15,17,29–34).<br />

The mechanisms underlying enhancement of PCR by many of these cosolvents are not<br />

well defined. It is suggested that some cosolvents, such as DMSO, formamide, glycerol,<br />

and polyethylene glycol, may affect the Tm of the primers, the thermal activity profile of<br />

Taq DNA polymerase, as well as the degree of product strand separation (18). Gelatin,<br />

bovine serum albumin, and nonionic detergents, such as Tween-20 and Triton X-100,<br />

are thought to stabilize DNA polymerases (18). TMAC is used to eliminate nonspecific<br />

priming (34). (NH 4 ) 2 SO 4 may increase the ionic strength of the reaction mixture, altering<br />

the denaturation and annealing temperatures of DNA, and may affect polymerase activity.<br />

Betaine has been shown to increase the thermostability of DNA polymerases, as well<br />

as to alter DNA stability such that GC-rich regions melt at temperatures more similar<br />

to AT-rich regions (29).<br />

It is important to carefully choose the appropriate cosolvents and correct concentrations<br />

to effectively improve PCR amplifications. Cosolvent concentrations should be no greater


96 Grunenwald<br />

than absolutely necessary for optimal amplification, as they may reduce DNA polymerase<br />

activity. For example, DMSO at a final concentration of 10% can reduce Taq DNA<br />

polymerase activity by up to 50% (15). In addition to improving standard PCR, multiplex<br />

PCR performance has also been shown to improve when using DMSO (34), Tween-20 and<br />

Triton X-100 (32), β-mercaptoethanol (33), TMAC (34), and betaine (35). Long PCR has<br />

been enhanced using glycerol, gelatin (17), DMSO (19), and betaine (36).<br />

9. The “hot start” technique enhances PCR specificity by eliminating the production of<br />

nonspecific products and primer-dimers during the initial steps of PCR (36). This is<br />

because even a brief incubation of a PCR mix at temperatures significantly below the Tm<br />

can result in primer-dimer formation and nonspecific priming. The purpose of a hot start<br />

is to withhold one of the critical components from the reaction until the temperature<br />

in the first cycle rises above the annealing temperature. There are various methods of<br />

performing a hot start. Manual hot start is performed by withholding one of the reaction<br />

components, such as the DNA polymerase or magnesium, and adding it only after the<br />

reaction temperature rises above 80°C during the first denaturation step. Wax-mediated<br />

hot start involves addition of a wax layer separating the component being withheld from<br />

the remainder of the reaction mix. During the temperature increase in the first denaturation<br />

step, the wax melts and the withheld component is mixed with the rest of the reaction<br />

components, starting the amplification reaction. The beads for wax layer can be made in<br />

the laboratory (37,38) or purchased commercially (Ampliwax PCR Gems, PerkinElmer).<br />

Hot start Taq DNA polymerase is constructed through the addition of an anti-Taq DNA<br />

polymerase antibody (TaqStart Antibody, Clontech). The antibody will prevent the<br />

DNA polymerase activity until the temperature rises during the initial denaturation step.<br />

The increased temperature dissociates and degrades the bound antibody, initiating PCR<br />

amplification. Hot start is commonly used for multiplex and long PCR amplifications.<br />

10. In addition to all of the PCR optimization strategies discussed above, there are also<br />

commercially available buffer systems for fast and easy PCR optimization. Companies,<br />

such as Boehringer Mannheim, Stratagene, Invitrogen, and Epicentre Technologies, offer<br />

various buffer systems for PCR optimization. These buffer systems can be divided into<br />

two categories. One category (e.g., PCR optimization kit from Boehringer Mannheim)<br />

contains a set of 16 buffers that combine different pH (8.3, 8.6, 8.9, and 9.2) and various<br />

concentrations of magnesium (1.0, 1.5, 2.0, and 3.5 mM). There are also four different<br />

cosolvents, DMSO, glycerol, gelatin, and (NH 4 ) 2 SO 4 , provided separately for additional<br />

optimization. Because the cosolvents are not premixed in the buffers, inclusion of these<br />

cosolvents will require a second set of optimization reactions. The other category of<br />

buffer system (e.g., PCR optimization kits from Epicentre Technologies) contains variable<br />

concentrations of magnesium (1.5, 2.5, and 3.5 mM) and a betaine-containing enhancer.<br />

Because all necessary reaction components, including the cosolvent (betaine), are premixed<br />

in this buffer system, only one set of optimization reactions are performed.<br />

11. The following conditions may lead to less than optimal PCR amplifications. The possible<br />

solutions for each condition are discussed.<br />

If little or not desired PCR product is detected:<br />

a. Too little DNA template is present in the reaction. Increase the amount of template<br />

DNA.<br />

b. The template DNA is damaged or degraded. Assure the purity and integrity of the DNA<br />

template by minimizing damage from nicking and shearing.<br />

c. Insufficient DNA polymerase is present in the reaction. Increase the DNA polymerase<br />

concentration in increments of 0.5 units per 100 µL of reaction.<br />

d. Insufficient number of cycles was performed. Increase cycle number by 5 to 10 cycles.<br />

e. Check for inhibitor(s) during template DNA preparation. Repurification of the DNA<br />

template may remove some inhibitors of PCR.


Optimization of PCRs 97<br />

f. Magnesium concentration is too low. Increase magnesium concentration in increments<br />

of 0.1 mM.<br />

g. The denaturation time is too long or too short. Adjust denaturation time in increments<br />

of 5 s.<br />

h. Add cosolvents that enhance PCR amplification.<br />

i. The denaturation temperature is too high or too low. Change denaturation temperature<br />

in increments of 1°C.<br />

j. The primer annealing temperature is too high. Lower annealing temperature in increments<br />

of 2°C.<br />

k. The primer extension period is too short. Increase extension time in increments of<br />

1 minute.<br />

l. Re-amplify dilutions (110 to 11000) of the first round of PCR amplification using<br />

nested primers.<br />

m. Perform hot start.<br />

n. Perform Touchdown (TD)/Stepdown (SD) PCR cycling program. TD or SD PCR uses<br />

a temperature cycling protocol that is performed at decreasing annealing temperatures.<br />

The cycling program begins at an annealing temperature a few degrees above the<br />

calculated Tm of the primers. This ensures that the first primer-template hybridization<br />

events involve only those sequences with the greatest specificity. The annealing<br />

temperature is decreased 1 to 4°C every other cycle to approx 10°C below the calculated<br />

Tm to permit exponential amplification (39). Here is an example of a typical TD/SD<br />

cycling program: initial denaturation at 94°C for 1 mi followed by 20 cycles of 10 s at<br />

92°C and 20 s at 70°C with an 0.5°C decrease of temperature per cycle, then another<br />

20 cycles of 10 s at 92°C and 30 s at 60°C with a 1-s extension per cycle and hold at<br />

4°C. Hot start must be used with these cycling programs.<br />

o. Review primer design and composition. Design new primers and try PCR again.<br />

If multiple product bands or smear is detected:<br />

a. Too much DNA template is present in the reactions. Decrease the amount of DNA<br />

template in the reaction mix.<br />

b. Annealing temperature is too low. Increase annealing temperature in increments of<br />

2°C.<br />

c. DNA polymerase concentration is too high. Decrease enzyme concentration in increments<br />

of 0.5 units per 100-µL reaction.<br />

d. Magnesium concentration is too high. Decrease the magnesium concentration in<br />

increments of 0.1 mM.<br />

e. Denaturation time is too short. Increase the denaturation time in increments of 5 s.<br />

f. Denaturation temperature is too low. Increase the denaturation time in increments<br />

of 1°C.<br />

g. Cycle number is too high. Reduce the cycle number by 5 to 10 cycle.<br />

h. Perform hot start.<br />

i. Alter concentrations of cosolvents.<br />

j. Perform TD/SD PCR.<br />

k. Extension time is too long. Reduce the extension time in increments of 1 min.<br />

l. Check for carry-over contamination. Set up PCR in a different area.<br />

m. Review primer design and composition. Design new primers and try PCR again.<br />

References<br />

1. Higuchi, R. (1989) Using PCR to engineer DNA, in PCR Technology: Principles and<br />

Applications for DNA Amplifications (Erlich, H. A., ed.), Stockton Press, Inc., New York,<br />

pp. 61–70.


98 Grunenwald<br />

2. Foord, O. S. and Rose, E. A. (1995) Long-distance PCR, in PCR Primer (Dieffenback,<br />

C. W. and Dveksler, G. S., ed.), Cold Spring Harbor Laboratory Press, Cold Spring, NY,<br />

pp. 63–77.<br />

3. Edwards, A., Civitello, A., Hammond, H. A., and Caskey, C. T. (1991) DNA typing and<br />

genetic mapping with trimeric and tetrameric tandem repeats. Am. J. Hum. Genet. 49,<br />

746–756.<br />

4. Edwards, A., Hammond, H. A., Jin, L., Caskey, C. T., and Chakroborty, R. (1992) Geneticvariation<br />

at five trimeric and tetrameric tandem repeat loci in four human population<br />

groups. Genomics 12, 241–253.<br />

5. Klimpton, C. P., Gill, P., Walton, A., Urquhart, A., Millican, E. S., and Adams, M. (1993)<br />

Automated DNA profiling employing multiplex amplification of short tandem repeat loci.<br />

PCR Methods Appl. 3, 13–21.<br />

6. Sambrook, J., Fritsch, E. F., and Maniatis, T. (1989) Molecular Cloning: A laboratory<br />

Manual. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, New York.<br />

7. Lindahl, T. and Nyberg, B. (1972) Rate of depurination of native deoxyribonucleic acid.<br />

<strong>Bio</strong>chemistry 11, 3610–3618.<br />

8. Rychlik, W. and Rhoads, R. E. (1989) A computer program for choosing optimal oligonucleotides<br />

for filter hybridization, sequencing and in vitro amplification of DNA. Nucleic<br />

Acids Res. 17, 8543–8551.<br />

9. Lowe, T. M. J., Sharefkin, J., Yang, S. Q., and Dieffenback, C. W. (1990) A computer<br />

program for selection of oligonucleotide primers for polymerase chain reaction. Nucleic<br />

Acids Res. 18, 1757–1761.<br />

10. O’Hara, P. J. and Venezia, D. (1991) PRIMGEN, a tool for designing primers from multiple<br />

alignments. CABIOS 7, 533–534.<br />

11. Montpetit, M. L., Cassol, S., Salas, T., and O’Shaughnessy, M. V. (1992) OLIGOSCAN:<br />

A computer program to assist in the design of PCR primers homologous to multiple DNA<br />

sequences. J. Virol. Methods 36, 119–128.<br />

12. Suggs, S. V., Hirose, T., Myake, D. H., Kawashima, M. J., <strong>John</strong>son, K. I., and Wallace, R. B.<br />

(1981) Using purified genes, ICN-UCLA Symp. Mol. Cell. <strong>Bio</strong>l. 23, 683–693.<br />

13. Breslauer, K. J., Ronald, F., Blocker, H., and Marky, L. A. (1986) Predicting DNA duplex<br />

stability from the base sequence. Proc. Natl. Acad. Sci. USA 83, 3746–3750.<br />

14. Freier, S. M., Kierzek, R., Jaeger, J. A., Sugimoto, N., Caruthers, M. H., Neilson, T.,<br />

et al. (1986) Improved free-energy parameters for predictions of RNA duplex stability.<br />

Proc. Natl. Acad. Sci. USA 83, 9373–9377.<br />

15. Innis, M. A. and Gelfand, D. H. (1990) Optimization of PCRs, in PCR Protocols: A Guide<br />

to Methods and Applications (Gelfand, D. H., Sninsky, J. J., Innis, M. A., and White, H.,<br />

eds.), Academic Press, San Diego, CA, pp. 3–12.<br />

16. Innis, M. A., Myambo, K. B., Gelfand, D. H., and Brow, M. A. D. (1988) DNA sequencing<br />

with Thermus aquaticus DNA polymerase and direct sequencing of polymerase chain<br />

reaction-amplified DNA. Proc. Natl. Acad. Sci. USA 85, 9436–9440.<br />

17. Ohler, L. and Rose, E. A. (1992) Optimization of long distance PCR using a transposonbased<br />

model system. PCR Methods Appl. 2, 51–59.<br />

18. Gelfand, D. H. (1989) Taq DNA polymerase, in PCR Technology: Principles and Applications<br />

for DNA Amplification (Erlich, H. A., ed.), Stockton Press, New York, pp. 17–22.<br />

19. Cheng, S. (1995) Longer PCR amplifications, in PCR Strategies (Innis, M. A., Gelfand, D.<br />

H., and Sninsky, J. J., eds.) Academic Press, San Diego, CA, pp. 313–324.<br />

20. Brock, T. D. and Freeze, H. (1969) Thermus aquaticus gene, a non-sporulating extreme<br />

thermophile. J. Bacteriol. 98, 289–297.<br />

21. Giebel, L. B. and Spritz, R. A. (1990) Site-directed mutagenesis using the double-stranded<br />

DNA fragment as a PCR primer. Nucleic Acids Res. 18, 4947.


Optimization of PCRs 99<br />

22. Saike, R. K., Gelfand, D. H., Stoffel, S., Scharf, S. J., Higuchi, R. Horu, G. T., Mullis,<br />

K. B., and Erlich, H. A. (1988) Primer-directed enzymatic amplification of DNA with a<br />

thermostable DNA polymerase. Science 239, 487– 491.<br />

23. Flaman, J.-M., Frebourg, T., Moreau, V., Charbonnier, R., Martin, C., Ishioka, C., et al.<br />

(1994). A rapid PCR fidelity assay. Nucleic Acids Res. 22, 3259–3260.<br />

24. Cline, J., Braman, J. C., and Hogrefe, H. H. (1996) PCR fidelity of Pfu DNA polymerase<br />

and other thermostable DNA polymerases. Nucleic Acids Res. 24, 3546–3551.<br />

25. Sardelli, A. D. (1993) Plateau effect-understanding PCR limitations, in Amplifications: A<br />

Forum for PCR Users. Perkin–Elmer Corp., Norwald, CT, pp. 1.<br />

26. Saiki, R. K. (1989) The design and optimization of the PCR, in PCR Technology (Erlich,<br />

H. A., ed.), Stockton Press, New York, pp. 7–16.<br />

27. Kim, H. S. and Smithies, O. (1988) Recombinant fragment assay for gene targeting based<br />

on the polymerase chain reaction. Nucleic Acids Res. 16, 8887–8903.<br />

28. McConlogue, L., Brow, M. D., and Innis, M. A. (1988) Structure-independent DNA<br />

amplification by PCR using 7-deaza-2′-deoxyguanosine. Nucleic Acids Res. 16, 9869.<br />

29. Mytelka, D. S. and Chamberlin, M. J. (1996) Analysis and suppression of DNA polymerase<br />

pauses associated with a trinucleotide consensus. Nucleic Acids Res. 24, 2774–2781.<br />

30. Pomp, D. and Medrano, J. F. (1991) Organic solvents as facilitators of polymerase chain<br />

reaction. <strong>Bio</strong>Techniques 10, 58–59.<br />

31. Newton, C. R. and Graham, A. (1994) PCR, <strong>Bio</strong>s Scientific, Publishers Ltd., Oxford.<br />

32. Levinson, G., Fields, R. A., Harton, G. L., Palmer, F. T., Maddleena, A., Fugger, E. F.,<br />

et al. (1992) Reliable gender screening for human preimplantation embryos, using multiple<br />

DNA target-sequences. Hum. Reprod. 7, 1304–1313.<br />

33. Chamberlain, J. S., Gibbs, R. A., Ranier, J. E., Nguyen, P. N., and Caskey, C. T. (1988)<br />

Deletion screening of the Duchenne muscular dystrophy locus via multiplex DNA<br />

amplification. Nucleic Acids Res. 16, 11,141–11,156.<br />

34. Uggozoli, L. and Wallace, B. (1992) Application of an allele-specific polymerase chain<br />

reaction to the direct determination of ABO blood group genotypes. Genomics 12,<br />

670–674.<br />

35. Henke, W., Herdel, K., Jung, K., Schnorr, D., and Loening, S. A. (1997) Betaine improves<br />

the PCR amplification of GC-rich DNA sequences. Nucleic Acids Res. 25, 3957–3958.<br />

36. Hengen, P. N. (1997) Optimizing multiplex and LA-PCR with betaine. Trends <strong>Bio</strong>chem.<br />

Sci. 22, 225–226.<br />

37. Bassam, B. J. and Caetano-Anolles, G. (1993) Automated “hot start” PCR using mineral<br />

oil and paraffin wax. <strong>Bio</strong>Techniques 14, 30–34.<br />

38. Wainwright, L. A. and Seifert, H. S. (1993) Paraffin beads can replace mineral oil as an<br />

evaporation barrier in PCR. <strong>Bio</strong>Techniques 14, 34–36.<br />

39. Rous, K. H. (1995) Optimization and troubleshooting in PCR, in PCR Primer (Diefferbach,<br />

C. W. and Dveksler, G. S. eds.), CSH Press, New York, pp. 53–62.


100 Grunenwald


Subcycling PCR 101<br />

21<br />

Subcycling PCR for Long-Distance Amplifications<br />

of Regions with High and Low Guanine–Cystine Content<br />

Amplification of the Intron 22 Inversion of the FVIII Gene<br />

David Stirling<br />

1. Introduction<br />

Hemophilia A is an X-linked disorder caused by mutations in the factor VIII gene.<br />

Around 50% of all patients with severe hemophilia A share a common mutation. This<br />

intron 22 inversion results from homologous recombination of a sequence within<br />

intron 22 of the factor VIII gene and identical sequence around 500 kb telomeric to<br />

the gene. Although this inversion could be detected by Southern blotting, the development<br />

of a long polymerase chain reaction (PCR) assay, and its subsequent improvement<br />

by subcycling PCR (S-PCR), greatly facilitated the provision of diagnostic<br />

services (1,2).<br />

In S-PCR, the annealing/elongation step is composed of subcycles of shuttling<br />

between a low and a high temperature. S-PCR produces consistent amplification of the<br />

various segments produced by wild-type, mutant, and carrier individuals. S-PCR is a<br />

robust variant of PCR, which may be of use in amplification of long segments in which<br />

the guanine–cytosine (GC) content varies among the segments, multiplex amplification<br />

of long segments, and multiplex amplification of short segments in which the GC<br />

content varies among the segments. We have also found it greatly improves the success<br />

of amplification from partially degraded templates.<br />

2. Materials<br />

Except where otherwise stated, all reagents are from Sigma Chemical Company,<br />

Poole, UK.<br />

1. Thermal cycler.<br />

2. Expand Long Template PCR System (Roche).<br />

3. Dimethyl sulfoxide.<br />

4. DNTP (Promega) Supplied separately (dATP, dCTP, dGTP, dTTP) at concentrations of<br />

100 mM. Use at 10 mM. Aliquot 10 µL of stock dNTP into labeled tubes containing 90 µL<br />

of sterile distilled water. Store at –20°C.<br />

5. 7-Deaza GTP (Roche).<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

101


102 Stirling<br />

Table 1<br />

Thermal Cycler Program for Subcycling PCR<br />

Step Temp/command Time (mm:ss)/repetitions<br />

11 95 2:00<br />

12 94 00:10<br />

13 63 00:05<br />

14 68 5:00<br />

15 Go to step 3 4 items<br />

16 Go to step 2 15 times<br />

17 94 00:10<br />

18 63 00:10<br />

19 68 6:00 (+ 10s per cycle)<br />

10 Go to step 8 4 times<br />

11 Go to step 7 15 times<br />

12 68 30:00<br />

13 End<br />

6. Oligonucleotides: INT22-P (5′-GCC CTG CCT GTC CAT TAC ACT GAT GAC ATT ATG<br />

CTG AC-3′); INT22-Q (5′-GGC CCT ACA ACC ATT CTG CCT TTC ACT TTC AGT<br />

GCA ATA-3′); INT22-A (5′-CAC AAG GGG GAA GAG TGT GAG GGT GTG GGA<br />

TAA GAA-3′); and INT22-B (5′-CCC CAA ACT ATA ACC AGC ACC TTG AAC TTC<br />

CCC TCT CAT A-3′). The oligonucleotides are diluted to 100 pmol/µL for storage.<br />

Subaliquot to minimize freeze-thaw cycles and store at –20°C.<br />

7. PCR Mastermix 1. For each reaction to be run, the following are included: 10× buffer 2<br />

(2.5 µmL); TAQ mix (0.94 µL); and water (5.69 µL).<br />

8. PCR Mastermix 2. For each reaction to be run, the following are included: 10 mM aATP<br />

(1.25 µL); 10 mM aTTP (1.25 µL); 10 mM aACP (1.25 µL); 10 mM aGTP (0.625 µL);<br />

10 mM deaza aGTP (0.625 µL); P/Q primer mix (5 µL); A (or B) primer mix (5 µL);<br />

100% DMSO (1.875 µL).<br />

3. Procedure<br />

1. Remove reagents from freezer and allow to thaw fully then mix thoroughly and briefly<br />

centrifuge before pipetting. Reagents to be thawed: buffer 2 from Expand Long Template<br />

PCR Kit, dNTPs, and deaza dGTP.<br />

2. Label 0.2 mL of flat-cap PCR tubes with the DNA number and primer grouping, if<br />

applicable (e.g., APQ).<br />

3. Using sterile pipet tips, pipet 0.5 µL of each DNA sample into the appropriately labeled<br />

tube. Ensure that the DNA is pipetted directly into the bottom the tube. If concentration of<br />

DNA is very low, 1 µL of DNA should be added. Place tubes on ice.<br />

4. Prepare 1 in 50 dilution of A oligo, 1 in 50 dilution of B oligo, and 1 in 25 dilution of<br />

P and Q oligo pairing. The number of samples being tested will determine the actual<br />

quantities required.<br />

5. Prepare PCR mastermix 1 and 2 for appropriate number of tests: Place both mastermixes<br />

on ice!<br />

6. For each reaction, aliquot 14.9 µL of Master Mix 2 into the labeled PCR tubes (which<br />

already contain the DNA template). Keep on ice.<br />

7. Immediately before amplification, add 9.2 µL of Master Mix 1. Spin briefly and proceed<br />

to PCR step immediately.


Subcycling PCR 103<br />

8. Place tubes in thermal cycler and start the program (see Table 1).<br />

9. When cycle is complete, remove PCR products and keep in fridge prior to electrophoresis.<br />

10. Dilute 4 µL of each PCR product into 6 µL of sterile distilled water. This initial dilution<br />

may need adjusting depending on electrophoresis results picture.<br />

11. Electrophorese 1 µL of this diluted DNA on an 0.8% agarose gel.<br />

12. The inversion is associated with a band of 11 kb; the normal allele is associated with a<br />

band of 12 kb. An upper 12-kb band only indicates the inversion is not present; a lower<br />

11-kb band only indicates the inversion is present in an affected male; and bands at both<br />

11 kb and 12 kb indicate a female carrier of the inversion.<br />

References<br />

1. Liu, et al. (1998) Single tube polymerase chain reaction for rapid diagnosis of the inversion<br />

hotspot mutation in haemophilia A (letter). Blood 92, 1458–1459.<br />

2. Liu and Sommer (1998) Sub-cycling PCR for multiplex long distance amplifications of<br />

regions with High and Low GC content: Application to the inversion hotspot in the FVIII<br />

gene. <strong>Bio</strong>Techniques 25, 1022–1028.


104 Stirling


Rapid Amplification of cDNA Ends 105<br />

22<br />

Rapid Amplification of cDNA Ends<br />

Xin Wang and W. Scott Young III<br />

1. Introduction<br />

The identification and isolation of full-length cDNAs can be a frustrating and timeconsuming<br />

experience, especially for genes with a low abundance of expression or with<br />

large transcripts. Traditionally, full-length cDNAs are obtained from cDNA libraries<br />

by hybridization with radioisotope-labeled probes. This labor-intensive and tedious<br />

procedure often produces incomplete sequences and sometimes includes intronic<br />

sequence. To obtain full-length cDNAs, investigators had to rescreen libraries with<br />

larger numbers of clones (or upstream probes), not always successfully. The combination<br />

of rapid amplification of cDNA ends (RACE) and long-distance polymerase chain<br />

reaction (PCR) with high fidelity makes it possible to obtain full-length cDNAs quickly<br />

without constructing or screening a cDNA library.<br />

The principle of RACE is simple and elegant: An anchor sequence is added to the<br />

end of the cDNA to be used as PCR primer binding template. A universal primer<br />

complementary to the added anchor template is coupled with a gene-specific primer<br />

(based on a single short known sequence within the mRNA of interest) in a PCR to<br />

amplify regions with unknown sequence. Several strategies have been developed to<br />

isolate full-length cDNA using this anchored PCR technology, each using a unique way<br />

to add the anchor sequence to the end of the cDNA. In the first generation of RACE,<br />

homopolymeric tails (G or A) are added to 3′ end of cDNA to be used as an anchor<br />

sequence using the enzyme terminal deoxynucleotidyl transferase (1). The second<br />

generation of RACE technique is based on the ability of T4 RNA ligase to ligate<br />

a single-stranded anchor sequence to the 3′ end of the first-strand cDNA (2). Both<br />

methods are difficult to optimize because of inefficient enzymatic reactions. The third<br />

generation of RACE uses T4 DNA ligase to add a double-stranded anchor sequence to<br />

both ends of double-stranded cDNAs (3), thus the resulting anchored cDNAs are suitable<br />

for both 5′- and 3′-RACE. A commercial kit called Marathon cDNA amplification<br />

kit has been built around this approach (Clontech Laboratories, Inc.).<br />

In this chapter, the application of this third-generation RACE method to the isolation<br />

of several pineal-specific cDNAs ranging from 1.4 to 8.0 kb in size is outlined. In<br />

one case, a full-length 2.0-kb cDNA for a pineal-specific cDNA, PG25, was obtained<br />

by 5′-RACE using gene-specific primers based on 260 base pairs of known sequence<br />

located in the 3′ terminus of the mRNA (4). In another case, two different versions of<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

105


106 Wang and Young<br />

full-length cDNAs for a pineal-specific gene, PG23, were obtained in a single 5′-RACE<br />

reaction because the antisense gene-specific primer used was derived from the common<br />

3′ portion of the mRNAs (277 bp known sequence from differential display PCR or<br />

DD-PCR, this chapter). This approach is suitable for cloning full-length cDNA quickly<br />

based on a short sequence <strong>info</strong>rmation from the 3′ end of mRNA, such as that obtained<br />

by the DD-PCR technique (5) or expressed sequence tags (6). PG10.2 is a gene (8 kb<br />

mRNA) expressed only in the pineal gland and the outer nuclear layer of the retina (7).<br />

Two 4-kb cDNA fragments were obtained by 5′-RACE and 3′-RACE using primers<br />

derived from only a 145-bp known sequence. Thus, this approach is suitable for cloning<br />

full-length cDNA based on a short known sequence located anywhere on the mRNA,<br />

whether derived from arbitrarily primed PCR (AP-PCR RNA fingerprinting technique<br />

(8) or suppression subtractive hybridization (SSH) technology (Clontech, ref. 9). In<br />

the case of PG10.2, a traditional 3′-RACE protocol was used to obtain the 3′ portion<br />

of unknown sequence. Then, an anchored cDNA pool constructed using an antisense<br />

gene-specific primer for reverse transcription was used to obtain the 5′ portion of<br />

unknown sequence. This approach is especially reliable and useful for isolating fulllength<br />

cDNAs for large transcripts (such as 8 kb for PG10.2) with low and restricted<br />

expression. It may take only a few weeks to go from identification of differentially<br />

expressed sequence tags (by DD-PCR, AP-PCR, or SSH technology) to full-length<br />

cDNA by this long-template PCR-based RACE. In the following sections, both<br />

5′-RACE and 3′-RACE protocols using the above model systems will be described.<br />

The efficiency of described RACE approaches are very satisfactory for obtaining<br />

full-length cDNAs quickly. The choice of method is dependent on the available<br />

resources and relevant experience. For some genes, the use of more than one method<br />

may be necessary. Another excellent approach for full-length cDNA cloning is using<br />

biotin-tailed oligonucleotide probes to capture full-length cDNA clones from a wellconstructed<br />

cDNA library (Invitrogen). A commercial kit called GeneTrapper cDNA<br />

Positive Selection System has been built around this capture technology (Invitrogen).<br />

2. Materials<br />

1. RNA from source tissue (see Note 1): pineal gland and retina tissues were dissected<br />

from Sprague–Dawley rats (male, 200–250 g, Taconic Farms). Total cellular RNA was<br />

isolated using TRIzol Reagent (Invitrogen) according to the manufacturer’s instructions.<br />

For preparation of double-stranded cDNA, poly (A)+ RNA was purified from 500 µg total<br />

RNA using a Poly (A) Quik mRNA isolation kit (Stratagene, La Jolla, CA).<br />

2. dNTP mixture: an aqueous solution of each dNTP (dGTP, dATP, dTTP, and dCTP) from<br />

any reputable vendor of molecular biology reagents.<br />

3. First-strand cDNA synthesis: SuperScript II (RNAse H-) reverse transcriptase (Invitrogen)<br />

was used with manufacturer supplied 5 X reverse transcription buffer (see Note 2).<br />

4. cDNA synthesis primer: this may be a universal oligo-dT primer (5′-CACTATAGGC<br />

CATCGAGGCC(T) 20 MN-3′) for 5′-RACE and/or 3′-RACE (see Subheading 3.2.1.) or an<br />

antisense gene-specific primer (5′-RACE only, see Subheading 3.4.2.) complementary to<br />

the rare mRNA of interest or located upstream on a large transcript (see Note 3).<br />

5. Second-strand cDNA synthesis: a 20× second-strand enzyme mixture, 5× second-strand<br />

buffer and T4 DNA polymerase supplied in the Marathon cDNA amplification kit (Clontech<br />

Laboratories, Inc.). Combining the following enzymes makes extra second-strand enzyme<br />

mixture: Escherichia coli DNA polymerase I, E. coli DNA ligase, and E. coli RNAse H<br />

(see Note 4).


Rapid Amplification of cDNA Ends 107<br />

6. Adaptor (anchor sequence) ligation: A specially designed partially double-stranded adaptor<br />

supplied in the Marathon cDNA amplification kit (Clontech Laboratories, Inc.).<br />

7. Oligonucleotide primers: gene-specific primers should be about 25 nucleotides long and<br />

around 50% guanine–cytosine. Primers with a melting temperature between 65 and 70°C<br />

give sufficient binding specificity (see Note 5). For some difficult genes, primers with<br />

70°C melting temperature or higher should be used in a touchdown PCR (see Note 6).<br />

Anchor primers (AP) complementary to the adaptor sequence (for 5′-RACE or 3′-RACE) or to<br />

the tailing sequence on the cDNA synthesis oligo-dT primer are coupled with GSP to amplify<br />

the unknown sequences flanked by the paired primer set. AP-1 (5′-CCATCCTAATACGACT<br />

CACTATAGGGC-3′) and AP-2 (nested within the AP-1, 5′-ACTCACTATAGGGCTC<br />

GAGCGGC-3′) are supplied in Clontech’s Marathon cDNA Amplification Kit.<br />

8. PCR machine: all experimental data presented in this paper were carried out on a<br />

thermocycler from MJ Research, Inc (see Note 7).<br />

9. DNA Polymerase: Expand PCR system (Boehringer-Mannheim). There are three buffer<br />

systems supplied by the manufacturer. Buffer 1 is sufficient for RACE PCR of expected<br />

product size of less than 10 kb.<br />

10. PCR fragment purification: QIAEX II from QIAGEN <strong>Bio</strong>science Corporation.<br />

11. TA cloning: pGEM-T vector system (Promega Corporation).<br />

12. Other general molecular laboratory equipment and reagents.<br />

13. All reagents, including enzymes, should be mixed briefly immediately before use.<br />

3. Methods<br />

3.1. The Relative Abundance of the Transcript in the Source Tissue<br />

1. A good understanding of expression profile is essential for the successful isolation of the<br />

full-length cDNA of interest. The efficiency of RACE PCR amplification largely depends<br />

on the relative abundance of the mRNA of interest in the poly (A)+ RNA sample extracted<br />

from the target tissue. RACE PCR should be performed on the tissue where the expression<br />

is most abundant. The higher the copy number of the mRNA in the cDNA pool, the better<br />

chance the full-length cDNA can be amplified to a critical mass visible on agarose gel<br />

(see Note 8).<br />

2. PG23 is a pineal-specific gene identified using DD-PCR (unpublished data). In this case,<br />

we had no choice but to use pineal as our target tissue for 5′-RACE. Northern blot analysis<br />

revealed a mRNA doublet about 2.0 kb and 2.4 kb for PG23 (Fig. 1A, lane 1). The size of<br />

mRNA serves as a useful guide in identifying the correct RACE bands (Fig. 1B, lanes 1<br />

and 2). In a similar fashion, another 2.0-kb full-length pineal-specific cDNA (PG25) was<br />

obtained using the RACE protocol outlined for PG23 (Fig. 3B in ref. 4).<br />

3. PG10.2 is a gene (8 kb mRNA) expressed only in the pineal gland and the outer nuclear<br />

layer of the retina (7). In this case, retina was used as target tissue source because PG10.2<br />

has a much higher expression level in retina than in pineal even though it is only expressed<br />

in the outer nuclear layer of the retina (Fig. 3 in ref. 7). This report will describe the<br />

application of a traditional 3′-RACE (see Subheading 3.3.) and a gene-specific 5′-RACE<br />

(see Subheading 3.4.) to the isolation of the large transcript of PG10.2 (8.0 kb). A general<br />

strategy is schematically depicted in Fig. 2.<br />

3.2. 5′-RACE Using a Universal Oligo-dT Primer for cDNA Synthesis<br />

(see Note 9)<br />

1. First-strand cDNA synthesis: 1 µg of high-quality poly (A)+ mRNA isolated from fresh<br />

rat pineal was reverse-transcribed to first-strand cDNA in a 10-µL reaction containing<br />

1 µL of 10× first-strand synthesis buffer, 1 µL of 0.1 M dithiothreitol, 1 µL of 10 mM


108 Wang and Young<br />

Fig. 1. (A) Northern blot analysis: 10 µg of total RNA were loaded in each lane; Pineal<br />

(lane 1), brain (lane 2), lung (lane 3), and kidney (lane 4). (B) RACE PCR: full-length<br />

cDNAs (1.4 kb and 2.0 kb) were obtained using a long template PCR-based RACE technique.<br />

Two primers, AGSP (5′- GGAACAGTCTGAGCTCTAAGCTAGG-3′, lane 1) and NAGSP<br />

(5′-CTAGAAGGATAAGTTCACGAGGGCC-3′, lane 2), were designed using a 277-bp<br />

sequence <strong>info</strong>rmation from a sequenced DD-PCR product (Fig. 2) and coupled with AP-1<br />

for 5′-RACE PCR.<br />

dNTP, 1 µL of 10 µM oligo-dT primer (5′-CACTATAGGCCATCGAGGCC(T)20MN-3′),<br />

and 1 µL of SuperScript II (RNase H-) reverse transcriptase (added last; see below). The<br />

reaction was performed in a thermocycler with the following parameters: 70°C for 10 min,<br />

42°C for 50 min, and 50°C for 15 min. SuperScript II RNase H- reverse transcriptase<br />

(1 µL) was added to the reaction mixture after a 5-min incubation at 42°C. The reaction<br />

mixture was placed on ice before the next step.<br />

2. Second-strand cDNA synthesis: The following components for second-strand cDNA<br />

synthesis were added to the above 10-µL first-strand reaction mixture: 48.4 µL of highquality<br />

H 2 O, 16 µL of 5× second-strand buffer, 1.6 µL of 10 mM dNTP mixture, and 4 µL<br />

of 20× second-strand enzyme cocktail. After mixing the contents briefly with gentle pipetting,<br />

the reaction was placed in the thermocycler for incubation at 16°C for 1.5 h. Then, 2 µL of<br />

T4 DNA polymerase was added and mixed well, and incubation continued at 16°C for<br />

another 45 min to create blunt end. Finally, 4 µL of the EDTA/Glycogen mixture (glycogen<br />

helps bring down the cDNA in a later precipitation) was added to stop the reaction.<br />

3. Purification of blunt-ended double-stranded cDNA: For efficient recovery of the relatively<br />

small amount of the cDNA, the general phenol/chloroform extraction was performed twice<br />

using 100 µL of phenolchloroformisoamyl alcohol (25241). The supernatant plus<br />

0.5 volume of 4 M ammonium acetate and 2.5 volumes of 100% ethanol were placed at –20°C<br />

for at least 1 h and centrifuged for 10 min at 4°C. The recovered DNA pellet was washed in<br />

1 mL of 75% ethanol and resuspended in 10-µL high-quality water after air-drying.<br />

4. Adaptor ligation: A partially double-stranded adaptor was ligated to both ends of the<br />

above-purified double-stranded cDNA. Ten microliters of ligation mixture containing<br />

5 µL of purified blunt-ended double-stranded cDNA, 2 µL of adaptor, 2 µL of 5× ligation


Rapid Amplification of cDNA Ends 109<br />

Fig. 2. Schematic representation of 5′- and 3′-RACE for the 8.0 kb cDNA of PG10.2. (A)<br />

The 145-bp sequence of PG10.2 is located in the middle of the 8-kb mRNA. (B) 3′-RACE.<br />

SGSP coupled with the anchor primer (complementary to the tail sequence of oligo-dT cDNA<br />

synthesis primer) for 3′-RACE. (C) 5′-RACE. AGSP2 coupled with the AP-1 anchor primer<br />

for 5′-RACE.<br />

buffer, and 1 µL of T4 DNA ligase were incubated at 16°C overnight. The mixture was<br />

heated at 70°C to inactivate the ligase.<br />

5. Primer design for PG23 5′-RACE: A short known sequence located at the 3′ end region<br />

of the gene is required for designing antisense gene-specific primers (AGSP) and a<br />

nested antisense gene-specific primer (NAGSP) for 5′-RACE. In the case of PG23, AGSP<br />

(5′- GGAACAGTCTGAGCTCTAAGCTAGG-3′) and NAGSP (5′-TCTAGAAGGATA-<br />

AGTTCACGAGGGCC-3′) were designed based on the 277-bp sequence <strong>info</strong>rmation just<br />

upstream of the poly (A) tail as shown in Fig. 3.<br />

6. PCR was performed in a 50-µL reaction containing the following components: 5 µL of<br />

10× PCR buffer 1, 5 µL of the 1:100 diluted adaptor-ligated cDNAs, 1 µL of 10 mM<br />

dNTP, 1 µL of 10 µM AP-1 linker primer, 1 µL of 10 µM GSP, and 0.75 µL of DNA<br />

polymerase mixture (added last). The DNA polymerase mixture was added to the reaction<br />

after denaturation at 94°C for 1 min to reduce nonspecific amplification (see Note 10).


110 Wang and Young<br />

Fig. 3. Expressed sequence tag of PG23. Sequence <strong>info</strong>rmation of 277 bp was derived<br />

from a cloned DNA fragment using the DD-PCR method (5). Double-strand sequences are shown<br />

for easy identification of primer’s location (underlined). A 10-nucleotide primer (5′-GATCT-<br />

GACTGC-3′) and an olio-dT primer (5′-TTTTTTTTTTTTCA-3′) were used in DD-PCR for<br />

the original cloning of PG23. AGSP (5′- GGAACAGTCTGAGCTCTAAGCTAGG-3′) and<br />

NAGSP (5′-TCTAGAAGGATAAGTTCACGAGGGCC-3′) were designed for the 5′-RACE of<br />

full-length cDNA of PG23.<br />

The sample was then subjected to 30 cycles of PCR on a thermocycler using the following<br />

parameters: 94°C for 30 s, 60°C for 30 s, and 68°C for 3 min. For RACE PCR of GC rich<br />

mRNAs, addition of some co-solvent may increase PCR efficiency (see Note 11).<br />

7. Agarose gel electrophoresis of 5′-RACE PCR products: The above AGSP and NAGSP<br />

were used in two independent RACE PCR. Then, 20 µL was loaded on a 1% agarose gel<br />

and electrophoresed in 1× TAE buffer at 75 volts for 2 h. Each of these antisense primers<br />

produced bands of expected size on ethidium bromide stained agarose gels when coupled<br />

with the AP-1 primer (Fig. 1B, lanes 1 and 2). The band produced by the NAGSP (Fig. 1B,<br />

lane 2) is slightly smaller compared with the one generated by AGSP (Fig. 1B, lane 1),<br />

confirming that the bands are true RACE products. When only AGSP, NAGSP, or AP-1<br />

were used in PCRs, these sharp bands disappeared (data not shown). The DNA bands of<br />

proper size were excised from agarose gel with a clean, sharp scalpel. DNA was extracted<br />

from the gel using the QIAEX II kit according to the manufacturer’s instruction. A 15-µL<br />

elution was saved for later use.<br />

8. The gel-purified DNA (5 µL) was used for cloning into the pGEM-T vector. Sequence<br />

<strong>info</strong>rmation for PG23 mRNAs of 1.4 kb and 2.0 kb was subsequently determined (data<br />

not shown).<br />

9. Nested second PCR: If the first PCR with AGSP and AP-1 is insufficient to produce a<br />

visible band(s) of expected size, a second round of PCR with NAGSP and AP-2 (nested<br />

within AP-1 primer) could be performed using 2.5 µL of 1:100 dilution of the first-PCR<br />

product as template. This second round of PCR usually yields defined fragments of<br />

expected size on ethidium bromide stained agarose gel. Sometimes no visible band


Rapid Amplification of cDNA Ends 111<br />

Fig. 4. 5′- and 3′-RACE of PG10.2. An antisense gene-specific primer 2 (AGSP2 of PG10.2,<br />

5′-GGCAGTTCATCCACACTCAGGTACCCAG-3′) and AP-1 primer amplified a faint but<br />

sharp band of about 3.5 for 5′-RACE of PG10.2 (lane 1). Either AGSP2 (lane 2) or AP-1<br />

(lane 3) alone failed to produce the band. A sense gene-specific primer (SGSP of PG10.2,<br />

5′-GAGTGAGAAGCAAAGTGCAAATGCC-3′) and an anchor primer (5′-CCAAGCTATT-<br />

TAGGTGACACTATAGGCCATCGAGGCC-3′, priming at the end of the newly synthesized<br />

second-strand cDNA) amplified a band of 4 kb for the 3′-RACE (lane 4).<br />

appeared even after a second round of PCR using nested primer set. This may due to nonoptimized<br />

RACE PCR conditions or to an extremely low level of expression of the gene.<br />

Other methods could be used to identify the correct RACE clones (see Note 12).<br />

3.3. 3′-RACE Using Tailed Oligo-dT Primer for cDNA Synthesis<br />

(see Note 9)<br />

1. Northern analysis revealed a mRNA of 8.0 kb in size for PG10.2. Only 145 bp of sequence<br />

<strong>info</strong>rmation for this DD-PCR fragment was available to be used as template for RACE<br />

primer design. After failing to find any positive cDNAs for PG10.2 in screens of two<br />

cDNA libraries, we suspected that this was probably caused by the relatively far upstream<br />

location of the 145 bp probe (Fig. 4 in ref. 7, underlined) and truncated clones in the cDNA<br />

libraries we examined. In order to obtain this potentially large 3′ portion of unknown<br />

sequence, we used a traditional 3′-RACE protocol outlined below to do 3′-RACE for<br />

PG10.2. In general, the extended sequence <strong>info</strong>rmation would provide a much broader<br />

region for 5′-RACE primer design.


112 Wang and Young<br />

2. First-strand cDNA synthesis: same as above (see Subheading 3.2.).<br />

3. Second-strand cDNA synthesis: same as above (see Subheading 3.2.). T4 DNA Polymerase<br />

was not used because blunt ending of cDNA was unnecessary in this case.<br />

4. Purification of double-stranded cDNA: same as above (see Subheading 3.2.).<br />

5. Primer design for PG10.2 3′-RACE: A sense gene-specific primer (SGSP of PG10.2,<br />

5′-GAGTGAGAAGCAAAGTGCAAATGCC-3′) was designed based on the known 145 bp<br />

sequence <strong>info</strong>rmation of PG10.2. An anchor primer (5′-CCAAGCTATTTAGGTGACAC-<br />

TATAGGCCATCGAGGCC-3′) was designed to prime at the end of the newly synthesized<br />

second-strand cDNA (Fig. 2). This end sequence on the newly synthesized second-strand<br />

cDNA is complementary to the 5′ tailing part of the oligo-dT primer used for first-strand<br />

cDNA synthesis.<br />

6. PCR was performed as described (see Subheading 3.2.) with SGSP of PG10.2 and the<br />

above-mentioned anchor primer. The PCR product was electrophoresed on a 1% agarose<br />

gel. A 4-kb fragment 3′ to the 145 bp known sequence was amplified very efficiently by<br />

this 3′-RACE PCR (Fig. 4, lane 4).<br />

7. The band was purified (as in Subheading 3.2.) and ligated into pGEM-T vector (as in<br />

Subheading 3.2.).<br />

3.4. 5′-RACE Using Gene-Specific Primer for cDNA Synthesis<br />

(see Note 9)<br />

1. To clone the full-length cDNA for the large transcript (8.0 kb mRNA) of PG10.2, a<br />

gene-specific RACE cDNA was generated and used as described below.<br />

2. First-strand cDNA synthesis was performed essentially as described (see Subheading 3.2.)<br />

except that an antisense gene-specific primer (AGSP1 of PG10.2, 5′-TTCAAGGGCCAGT-<br />

CAGGCCGTAGGTCACAGACACTTTGAC-3′) based on 145-bp sequence <strong>info</strong>rmation<br />

for PG10.2 was used for first-strand cDNA synthesis. The cDNA generated by this cDNA<br />

synthesis primer (GSP) is only suitable for 5′-RACE of this specific gene.<br />

3. Second-strand cDNA synthesis: same as above (see Subheading 3.2).<br />

4. Purification of double-stranded cDNA: same as above (see Subheading 3.2).<br />

5. Adaptor ligation: same as above (see Subheading 3.2.).<br />

6. Primer design for PG10.2 5′-RACE: An antisense gene-specific primer2 (AGSP2,<br />

5′- GGCAGTTCATCCACACTCAGGTACCCAG-3′) based on 145 bp sequence <strong>info</strong>rmation<br />

was designed for 5′-RACE of PG10.2. Notably, this primer was designed to be<br />

upstream of the AGSP1 used for cDNA synthesis (see Subheading 3.4.). In this design,<br />

AGSP2 is a nested primer relative to AGSP1 (Fig. 2).<br />

7. PCR was performed as described (see Subheading 3.2). Agarose gel electrophoresis<br />

showed that the AGSP2 of PG10.2 coupled with AP1-linker primer (Clontech, complementary<br />

to the sequence of the ligated-adaptor) amplified a unique band of about 3.5 kb<br />

in size upstream of the 145-bp known sequence of the PG10.2 fragment (Fig. 2, lane1).<br />

When only the AGSP2 of PG10.2 or AP-1 was used in PCRs, this faint but sharp band<br />

disappeared (Fig. 4, lanes 2 and 3).<br />

8. The band was purified (see Subheading 3.2.) and ligated into pGEM-T vector (as in<br />

Subheading 3.2.).<br />

9. The 4368-bp sequence, including the entire 5′-RACE product and the 5′-portion of the<br />

3′-RACE product, was determined (Fig. 4 in ref. 7).<br />

4. Notes<br />

1. Template purity (free of DNA) and integrity (minimum degradation) are critical for<br />

effective isolation of full-length cDNA of interest. One should place the dissected<br />

tissue immediately on dry ice before isolating total and poly (A)+ RNAs. All standard


Rapid Amplification of cDNA Ends 113<br />

precautions for handling RNA should be followed carefully (10). TRIzol Reagent<br />

(Invitrogen) performs well with small quantities of tissue. Total RNA has been successfully<br />

extracted from punches of rat supraoptic and paraventricular nuclei and used to clone a<br />

gene preferentially expressed in the supraoptic and paraventricular nuclei of the brain<br />

by DD-PCR (unpublished data). High-quality poly (A)+ mRNA can be obtained from<br />

companies such as Clontech Laboratories, Inc.<br />

2. Reverse transcriptase and related buffers supplied in the Marathon cDNA amplification<br />

(Clontech) may be used for the first-strand synthesis to meet most full-length cDNA<br />

cloning needs. SuperScript II (RNAse H-) reverse transcriptase (Invitrogen) is a preferred<br />

reverse transcriptase to generate longer first-strand cDNAs.<br />

3. The oligo-dT primer (5′-TTCTAGAATTCAGCGGCCGC(T) 30 MN-3′) supplied in the<br />

Marathon cDNA amplification kit (Clontech) is suitable as a universal cDNA synthesis primer.<br />

The two degenerate nucleotides (where M=A, G or C; N=A, G, T or C) place the oligo-dT<br />

primer at the beginning of the poly(A) tail and thus eliminate 3′-heterogeneity (11).<br />

4. Combine 25 µL of E. coli DNA polymerase I (10.0 units/µL, Invitrogen), 3 µL of E. coli<br />

DNA ligase (10.0 units/µL, Invitrogen), 3 µL of E. coli RNase H (2.1 units/µL, Invitrogen),<br />

and 10 µL of high-quality H 2 O to make 20× second-strand enzyme mixture. 5× secondstrand<br />

buffer contains 500 mM KCL, 50 mM ammonium sulfate, 25 mM MgCL 2 , 0.75 mM<br />

beta-NAD, 100 mM Tris (pH 7.5), and 0.25 mg/mL bovine serum albumin (Clontech’s<br />

Marathon cDNA Amplification Kit).<br />

5. The primer’s melting temperature is an important parameter that greatly influences PCR<br />

specificity by reducing nonspecific priming events. High-melting point primers with<br />

melting temperatures between 65 and 70°C allow the use of higher annealing temperatures<br />

to enhance reaction specificity. The estimated melting temperature [(G+C) × 4 + (A+T) × 2]<br />

are not exact under PCR conditions but can be used as a starting point. Sense primers are<br />

used for 3′-RACE whereas anti-sense primers are used for 5′-RACE.<br />

6. Touchdown PCR may help eliminate extraneous bands and increase yield. The first round<br />

of touchdown PCR has an annealing temperature 5 to 10°C higher than what is usually<br />

used. In each subsequent round, the annealing temperature is dropped a degree until the<br />

standard annealing temperature is reached.<br />

7. GeneAmp PCR System 9600 (Applied <strong>Bio</strong>systems) gives more satisfactory amplification<br />

efficiency, especially for low abundant genes. Applied <strong>Bio</strong>systems GeneAmp 0.5-ml PCR<br />

tubes are preferred PCR tubes for critical PCR amplification.<br />

8. Northern analysis is the most preferred method to provide an expression profile and<br />

size estimation of the mRNA. If northern data is not readily available, a quick reversetranscription<br />

PCR survey using two GSPs (to produce a sizable PCR fragment) could be<br />

used to identify the best tissue source for the actual RACE PCR.<br />

9. The 5′-RACE outlined in Subheading 3.2. is essentially as described in the Marathon<br />

cDNA amplification kit (Clontech Laboratories, Inc.) and the Expand PCR system<br />

(Boehringer Mannheim <strong>Bio</strong>chemicals) except an oligo-dT primer (5′- CACTATAGGCCATC<br />

GAGGCC(T) 20 MN-3′) was used for first-strand cDNA synthesis. The adaptor-ligated<br />

cDNA pool is essentially an uncloned cDNA library, which can be used to isolate fulllength<br />

cDNAs for many different genes using gene-specific primers. These two-sided<br />

(5′ and 3′) anchored cDNAs permit 5′- and 3′-RACE PCR to be performed with the same<br />

cDNA pool. If the known sequence <strong>info</strong>rmation is far upstream from the 3′ end of the<br />

mRNA (poly A tail), especially for the large transcript of the gene, the 3′-RACE protocol<br />

described in Subheading 3.3. is a preferred strategy to obtain this potentially long (often<br />

untranslated) region quickly and reliably. Although the ligation of the double-stranded<br />

adaptor using T4 DNA ligase is much more efficient compared to the inefficient tailing or<br />

the single-stranded anchor ligation, the adaptor actually tags only a portion of cDNAs. The


114 Wang and Young<br />

3′-RACE outlined in Subheading 3.3. is more efficient because it relies on the sequence<br />

introduced by the tail sequence 5′ of the oligo-dT sequence as the template for anchor<br />

primer binding. The cDNA pool always provides more template for this kind 3′-RACE<br />

than the one available from ligated adaptors, hence giving more robust 3′-RACE efficiency.<br />

In general, the newly extended sequence <strong>info</strong>rmation provides a larger region for design<br />

of 5′-RACE primers. This is especially important if the available sequence <strong>info</strong>rmation for<br />

primer design is very limited. The gene-specific 5′-RACE protocol outlined in Subheading<br />

3.4. is useful for obtaining full-length cDNAs of large transcripts or those expressed<br />

at low levels. The cDNA produced by this protocol is only suitable for RACE PCR of<br />

unknown sequence upstream of the gene-specific primer that was used for first-strand<br />

cDNA synthesis.<br />

10. Hot start PCR. The DNA polymerase is withheld from the reaction until the temperature<br />

of reaction tube is above the annealing temperature. This hot-start PCR is used to improve<br />

the specificity and sensitivity of the RACE PCR.<br />

11. Co-solvent addition. For some primer/template systems such as GC-rich sequences, the<br />

addition of glycerol and/or DMSO to a final concentration of 5% has been found to<br />

enhance PCR yield and/or specificity.<br />

12. The RACE PCR product may be cloned into pGEM-T vector even if agarose gel electrophoresis<br />

fails to show a visible band(s). One dozen to one hundred colonies could be<br />

screened for positive clones in colony hybridization using oligodeoxynucleotide probes<br />

downstream from the RACE primer used for RACE PCR.<br />

Acknowledgments<br />

All experimental data presented in this chapter were performed in the former<br />

Laboratory of Cell <strong>Bio</strong>logy, National Institute of Mental Health/National Institutes<br />

of Health. We would like to acknowledge Dr. Michael J. Brownstein for invaluable<br />

advice and generous support. We thank Dr. Herman E. Brockman for critical reading<br />

of the manuscript.<br />

References<br />

1. Frohman, M. A., Dush, M. K., and Martin, G. R. (1988) Rapid production of full-length<br />

cDNAs from rare transcripts: amplification using a single gene-specific oligonucleotide<br />

primer. Proc. Natl. Acad. Sci. USA 85, 8998–9002.<br />

2. Edwards, J. B. D. M., Delort, J., and Mallet, J. (1991) Oligodeoxyribonucleotide ligation<br />

to single-stranded cDNAs: A new tool for cloning 5′ ends of mRNAs and for constructing<br />

cDNA libraries by in vitro amplification. Nucleic Acids Res. 19, 5227–5232.<br />

3. Chenchik, A., Diatchenko, L., Moqadam, F., Tarabykin, V., Lukyanov, S., and Siebert,<br />

P. D. (1996) Full-length cDNA cloning and determination of mRNA 5′ and 3′ ends by<br />

amplification of adaptor-ligated cDNA. <strong>Bio</strong>Techniques 21, 526–532.<br />

4. Wang, X., Brownstein, M. J., and Young, W.S., III (1997) PG25, a pineal-specific cDNA,<br />

cloned by differential display PCR (DDPCR) and rapid amplification of cDNA ends<br />

(RACE). J. Neurosci. Methods 73, 187–191.<br />

5. Liang, P. and Pardee, A. B. (1992) Differential display of eukaryotic messenger RNA by<br />

means of polymerase chain reaction. Science 257, 967–971.<br />

6. Sikela, J. M. and Auffray, C. (1993) Finding new genes faster than ever. Nat. Genet. 3,<br />

189–191.<br />

7. Wang, X., Brownstein, M. J., and Young, W. S., III (1996) Sequence analysis of PG10.2,<br />

a gene expressed in the pineal gland and the outer nuclear layer of the retina. Mol. Brain<br />

Res. 41, 269–278.


Rapid Amplification of cDNA Ends 115<br />

8. McClelland, M., Ralph, D., Cheng, R., and Welsh, J. (1994) Interactions among regulators<br />

of RNA abundance characterized using RNA fingerprinting by arbitrarily primed PCR.<br />

Nucleic Acids Res. 22, 4419– 4431.<br />

9. Diatchenko, L., Lau, Y.-F. C., Campbell, A. P., Chenchik, A., Moqadam, F., Huang, B., et<br />

al. (1996) Suppression subtractive hybridization: A method for generating differentially<br />

regulated or tissue-specific cDNA probes and libraries. Proc. Natl. Acad. Sci. USA 93,<br />

6025–6030.<br />

10. Sambrook, J., Fritsch, E. F., and Maniatis, T. (1989) Molecular Cloning: A Laboratory<br />

Manual, 2nd Ed. Cold Spring Harbor Laboratory, Cold Spring Harbor, New York.<br />

11. Borson, N. D., Salo, W. L., and Drewes, L. R. (1992) A lock-docking oligo(dT) primer for<br />

5′ and 3′ RACE PCR. PCR Methods Appl. 2, 144–148.


116 Wang and Young


RAPD Fingerprinting 117<br />

23<br />

Randomly Amplified Polymorphic DNA Fingerprinting<br />

The Basics<br />

Ranil S. Dassanayake and Lakshman P. Samaranayake<br />

1. Introduction<br />

The study of genetic polymorphism among diverse populations of organisms is<br />

a complex task. However, this can be accomplished by using newer tools, such as<br />

randomly amplified polymorphic DNA (RAPD). RAPD is a polymerase chain reaction<br />

(PCR) technique that relies on the generation of amplification products for a given<br />

nucleic acid using an amplification-based scanning technique driven by arbitrary<br />

priming oligonucleotides. The result is the generation of amplification products<br />

(amplicons) that represent a multiplicity of anonymous sites that are characteristic of<br />

the studied genome (Fig. 1A,B).<br />

In RAPD, the first few cycles are performed at a low stringency, which ensures the<br />

generation of products by priming with mismatches between the primer and the template.<br />

The subsequent PCR cycles are performed at a higher stringency (see Note 1),<br />

yielding products that have ends complementary to the primer. The amplified region<br />

consists of unstructured, hypervariable, mostly noncoding sequences that vary in length<br />

from one species to another. As the arbitrary priming depends upon the complimentary<br />

regions in the DNA template, differences in these regions lead to uniquely characteristic<br />

fingerprinting patterns (1). These differences permit any organism to be characterized<br />

at the species or the strain level. However, it is noteworthy that the clarity of the species<br />

or strain discrimination, fingerprint complexity, and detection of DNA polymorphism<br />

are dependent on the primer that is selected for the RAPD assay.<br />

A single primer approx 10 bp in size (40–70% guanine–cytosine content) is generally<br />

used in PCR fingerprinting. For the amplification of the target region, the distance<br />

between priming regions has to be not more than 3 to 4 Kb. Specificity in RAPD is<br />

defined as the ability to produce “consensus” fingerprints on multiple occasions as<br />

a result of a multiplicity of targeted arbitrary sites. To achieve such consensus and<br />

“stringent” fingerprints in arbitrary priming protocols, a decrease in primer length (2)<br />

or blocking of the interactions between amplicons by incorporating a mini-hairpin at<br />

the 5′ terminus of the oligonucleotide is required (3).<br />

There are a number of limitations in the routine use of RAPD profiles for detailed<br />

evaluation of the genomes of organisms. The necessarily short primers require stringent<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

117


118 Dassanayake and Samaranayake<br />

conditions for reproducible PCR (4), because variable or absent PCR products may<br />

result depending on the purity, quantity, or the quality of the DNA templates (5).<br />

The reproducibility of the RAPD technique can also be affected by minor changes in<br />

methodological aspects, such as the differences in the primer-to-template concentration<br />

ratio (see Note 2), variations of primer annealing temperatures, the cation concentration<br />

of PCR buffer, and magnesium ions in the reaction mixture (6, see Note 3). These<br />

parameters can dramatically affect the presence of low-intensity bands as well as the<br />

position and intensity of high-intensity bands. Moreover, some have reported that<br />

different lots of Taq polymerases (see Note 4) and the brand of the thermocycler used<br />

(see Note 5) can also affect the RAPD patterns, especially the low-intensity bands (7).<br />

Furthermore, in some situations, the bands that show equal electrophoretic mobility<br />

may not be homologes, and missing bands may not necessarily reflect homology<br />

because they can be lost by nucleotide substitutions in either the PCR priming sites<br />

or by length mutations. These complications of identity can be resolved either by<br />

sequencing the homologous bands or using a band-specific probe. Such strategies have<br />

been used in several recent studies (8–10).<br />

Our studies are mostly focused on the genomic analysis of the human fungal<br />

pathogen Candida albicans, and the following protocol is based on this experience.<br />

However, the principles guiding RAPD analyses are similar, and the following methods<br />

with minor variations would be applicable in general for many other organisms.<br />

2. Materials<br />

1. Thermocycler.<br />

2. Agarose gel electrophoresis apparatus.<br />

3. Ultraviolet transilluminator, Power supply (200 V and 150 mA).<br />

4. Spectrophometer for determining DNA concentration.<br />

5. Photographic unit that can capture ethidium bromide stained gel (e.g., Polaroid camera<br />

or digital image capture system, such as a CCD camera, a computer, and image analysis<br />

software for clear resolution of the gel image).<br />

6. 10× PCR buffer: 100 mM Tris-HCl, 500 mM KCl, 15 mM MgCl 2 , 0.01% (w/v) gelatin,<br />

pH 8.3.<br />

7. 25 mM Magnesium chloride.<br />

8. High-quality sterile, deionized water (more than 10 megaohm/cm) must be used for the<br />

preparation of all reagents and premixes.<br />

9. Deoxynucleotide triphosphate stock solution (dNTP): 2 mM each of dGTP, dATP, dCTP,<br />

and dTTP. Ready-made dNTP (100 mM) solution (obtainable from Sigma, Pharmacia,<br />

Promega, etc.). Make aliquots, preferably 10 mM, and store in –20°C. If dNTP solutions<br />

are made from dry reagents, the pH of the solution should be adjusted to 7.5 with 0.1 M<br />

Tris or 0.1 M NaOH using a pH meter or strip of pH paper.<br />

10. Primers: Lyophilized primers (5′GCGATCCCCA3) should be prepared at 100 µM and 10 µM<br />

concentration with deionized water; Store at –20°C.<br />

11. Taq DNA polymerases: Taq DNA polymerase (5 U/µL; Sigma), Ampli Taq (5 U/µL;<br />

PerkinElmer), Stoffel fragment of Taq DNA polymerase (10 U/µL; Perkin–Elmer), or any<br />

other high-quality Taq polymerase is preferred.<br />

12. Template DNA: 10 to 25 ng/µL stock solution containing good-quality, protein-free,<br />

nonsheared DNA; DNA can be resuspended in high-quality sterile, deionized water or TE<br />

(Tris-EDTA) pH 8.0. RNase (20 ng per 1 ng of DNA) treatment can increase amplification<br />

several-fold. RNase added should be DNase-free.<br />

13. Sterile mineral oil.


RAPD Fingerprinting 119<br />

14. GeneAmp ® Thin-Walled Reaction Tubes (0.6 mL), designed for optimal fit in the DNA<br />

thermal cycler or DNA cycler 480 sample block.<br />

15. Agarose (Sigma or similar product).<br />

16. Ethidium bromide to visualize DNA<br />

17. TBE buffer: 1 M Tris, 0.9 M boric acid, 0.01 M EDTA, pH 8.3.<br />

18. Loading dye (6× concentration: 0.25% bromophenol blue, 0.25% xylene cyanol FF, and<br />

40% (w/v) sucrose in water, keep at 4°C.<br />

19. Electrophoretic size standards: PCR marker, 50–2000 bp, 123-bp DNA ladder, Supercoiled<br />

DNA ladder [2–16 kb], Sigma.<br />

3. Methods<br />

Programming of the thermal cycler for RAPD analysis is described first, followed<br />

by the PCR conditions used for the reaction.<br />

3.1. RAPD PCR Programming<br />

Program the thermal cycler (PerkinElmer model 480, which is a thermocycler with an<br />

average ramp speed) to denature for 1 min at 94°C, anneal for 2 min at 27°C (see Note 1),<br />

and allow primer extension for 2 min at 72°C for the first five cycles. Then, program<br />

for 45 cycles of 1 min denaturation at 94°C, 2 min of annealing at 32°C (see Note<br />

1), and 2 min primer extension at 72°C. Set the final extension period for 15 min<br />

at 72°C. With a thermocycler with a faster ramp speed, like the PerkinElmer model<br />

9600, a shorter protocol can be used, such as 45 cycles of 15 s at 94°C, 30 s at 27°C,<br />

and 1 min at 72°C.<br />

3.2. Setting Up the PCR for RAPD Analysis<br />

The following protocol is particularly suitable for RAPD analysis of C. albicans.<br />

Minor modifications of this method are useful for RAPD analysis of other Candida species.<br />

The components and procedure of the RAPD reaction mixture are listed as follows:<br />

1. Thaw the PCR buffer, MgCl 2 solution, dNTPs, and primer solutions on ice and mix<br />

properly before use.<br />

2. Prepare the PCR master mix in GeneAmp Thin-Walled Reaction Tubes containing<br />

approximately 50 ng of C. albicans genomic DNA (see Note 6), 5 µL of 10× PCR standard<br />

buffer, 200 µM dNTPs, 5 µL of 25 mM MgCl 2 (see Note 3), 1 µM of primer, 1.5 U Taq<br />

polymerase (Sigma, see Note 4), and make up to 50 µL of final reaction volume with<br />

double distilled deionized sterile water.<br />

3. Mix these reagents well by vortexing and then overlay with mineral oil (40 µL) to prevent<br />

evaporation and internal condensation. Spin for a short period before amplification until a<br />

smooth interface appears between the aqueous and the mineral oil layer.<br />

4. Preparation of the agarose gel is as follows: Separate the amplified products in 2% agarose<br />

gel because the molecular weight of the resultant amplicons are in the range of 0.3 to 4.2 kb<br />

(agarose has to be dissolved in the same electrophoresis running buffer, generally, 1× TBE.<br />

Include 0.5 µg/mL ethidium bromide during the preparation of the gel).<br />

5. Sample preparation for gel loading is as follows: Retrieve amplified products by adding<br />

150 µL of chloroform and pipetting out the aqueous droplet or by placing the reaction<br />

mixture over a piece of parafilm. Then, mix well the retrieved aqueous amplification<br />

mixture (12.5 µL) with the loading buffer (1.5 µL) and load into wells in an agarose<br />

gel together with an appropriate DNA size marker (ranging in size from 50 to 6 Kb) in<br />

a standard manner.


120 Dassanayake and Samaranayake<br />

5. Electrophorese amplified products under 5 to 10 V/cm for 2 to 2.5 h. When the bromophenol<br />

blue travels three quarters of the length of the gel, visualize the gel under ultraviolet<br />

light and photograph.<br />

4. Notes<br />

1. Annealing temperature: This is an important parameter that needs optimization in RAPD<br />

and depends on primer length and sequence. The melting temperature (Tm) of a primer<br />

is proportional to both its length and the G + C content. Tm for primer template can be<br />

determined using the formula given below.<br />

Tm = [(number of A+T) × 2°C + (number of G + C) × 4°C]<br />

However, optimal annealing temperature for a primer should be adjusted empirically.<br />

Generally, in RAPD, the first few cycles are performed at a low annealing temperature<br />

(`5°C below the calculated Tm) and subsequent cycles are performed at a high annealing<br />

temperature (`5°C above the calculated Tm).<br />

2. Primer-template ratio: This is one of the important variables in the amplification reaction.<br />

Generally, moderate primer:template mass ratios ranging from 0.5 to 5000 are used.<br />

3. Ionic composition: The concentration of ionic components is critical for RAPD. Of<br />

these, magnesium is important because different thermostable polymerases have different<br />

affinities for magnesium. Generally, the higher the concentration of the magnesium ions,<br />

the lower is the specificity, and vice versa. In our hands reproducible fingerprints for<br />

C. albicans isolates were obtained with magnesium ion concentrations of 2.5 mM. It is noteworthy<br />

that when DNA is dissolved in TE buffer, the magnesium ion concentration has to be<br />

increased (~3 mM) to obtain reproducible patterns. This is probably caused by the chelation<br />

of magnesium ions by EDTA, which lowers the effective ionic concentration in the reaction<br />

mixture. Further, Weaver et al. (11) reported that excess primer and template DNA could<br />

also modulate the activity of magnesium by sequestrating free magnesium ions and thus<br />

dampening the amplification reaction. The dNTP concentration also has a direct effect<br />

on the magnesium ion concentration in the reaction mixture as a result of the interaction<br />

between the mononucleotide and magnesium. Thus, a higher concentration of magnesium<br />

ions is necessary for amplifications with a higher concentration of dNTPs (12). On the<br />

contrary, high magnesium ion concentrations can lead to primer–template mismatching<br />

and thus decrease amplification stringency. Furthermore, magnesium ions can tightly bond<br />

with the sugar backbone of nucleotides and nucleic acids and therefore variation in the<br />

magnesium concentration has strong and complex effects on nucleic acid interactions.<br />

4. DNA polymerase: The activity of polymerases is highly variable (13) and therefore, subtle<br />

differences in the specificity of these enzymes can influence the fingerprint profiles, and the<br />

multiplex ratio (14,15). The polymerase activity is regulated to a great degree by the buffer<br />

components and, thus, a recommended buffer has to be used for a particular polymerase.<br />

We have observed that the variations in the combinations of buffers and polymerases lead<br />

to inadequately resolved and incomplete fingerprints. Highly variable results are obtained<br />

in particular when different eubacterial DNA polymerases are used in the RAPD technique.<br />

On the contrary, Thermal aquatics Stoffel fragment is a truncated DNA polymerase that<br />

has wide magnesium tolerance and thermal stability and produces well-defined low<br />

molecular weight products (less than 500 bp) in general. RAPD with truncated DNA<br />

polymerases are known to produce good yields. Generally, Taq polymerase concentrations<br />

of 1 to 1.25 U/50 µL-reaction is used in RAPD. However, the reaction mixture of more<br />

than 2 U/µL can generate nonspecific products (16).<br />

5. Thermal cycling parameters: These are of critical importance for optimization of the<br />

RAPD reaction (17). Thermal cycling parameters include template denaturation and


RAPD Fingerprinting 121<br />

Fig. 1. RAPD fingerprinting of 15 clinical Candida parapsilosis isolates (P1 to P10 superficial<br />

and P11 to P15 systemic) with primers RSD12 (5′ GCA TAT CAA TAA GCG CAG GAA<br />

AAG 3′) (A) and RSD6 (5′ GCG ATC CCC A 3′) (B) obtained after electrophoretic separation<br />

on 1.2% agarose gel. M, PCR marker (Sigma). Sizes of bands indicate the number of base<br />

pairs (18).<br />

annealing temperatures, cycle number and time duration of denaturation, annealing and<br />

primer extension period, and the type of thermal cycler used (4). Different varieties of<br />

thermocyclers, with intrinsic inhomogeneneities in rates of cooling and heating, can elicit<br />

incongruent fingerprints despite identical program settings and or reaction components.<br />

Similarly, subtle changes in the temperature within the same heat block can alter the<br />

mode of amplification. Annealing temperature also impacts the quality of the fingerprints<br />

produced and their reproducibility. It is noteworthy that the lifetime of a polymerizing agent<br />

can be extended considerably by reducing the duration of the denaturation temperature.<br />

6. Template: It is well known that the purity of the template is critical for producing<br />

reproducible fingerprinting patterns; thus, DNA devoid of proteins and RNA must be used<br />

at all times. For instance, impurities, such as phenol remnants in DNA, can be removed by<br />

repeated washing in 70% ethanol. Also DNA isolation methods that lead to degradation<br />

of DNA and inhibit the activity of DNA polymerase should be avoided. Furthermore,<br />

the RAPD technique cannot be reproducibly used to amplify DNA beyond a minimal<br />

threshold (less than 5 ng) or at the other extreme, a very high concentration of template<br />

DNA (higher than 1 µg). Such attempts invariably lead to production of either “smears”<br />

or poor resolution of the amplicons. In general, 50 to 100 ng of DNA can be used for<br />

50 µL of PCR reaction mixture (17).


122 Dassanayake and Samaranayake<br />

References<br />

1. Williams, J. G., Kubelik, A. R., Livak, K. J., Rafalski, J. A., and Tingey, S. V. (1990)<br />

DNA polymorphisms amplified by arbitrary primers are useful as genetic markers. Nucleic<br />

Acids Res. 18, 6531–6535.<br />

2. Caetano-Anolles, G., Bassam, B. J., and Gresshoff, P. M. (1992). Primer-template interactions<br />

during DNA amplification fingerprinting with single arbitrary oligonucleotides. Mol.<br />

Gene Genet. 235, 157–165.<br />

3. Caetano-Anolles, G. and Gresshoff, P. M. (1994). DNA amplification fingerprinting using<br />

arbitrary mini-hairpin oligonucleotide primers. <strong>Bio</strong>technology (NY) 12, 619–623.<br />

4. Penner, G. A., Bush, A., Wise, R., Kim, W., Domier, L., Kasha, K., et al. (1993) Reproducibility<br />

of random amplified polymorphic DNA (RAPD) analysis among laboratories. PCR<br />

Methods Appl. 2, 341–345.<br />

5. Kubelik, A. R. and Szabo, L. J. (1995). High-GC primers are useful in RAPD analysis of<br />

fungi. Curr. Genet. 28, 384–389.<br />

6. Ellsworth, D. L., Rittenhouse, K. D., and Honeycutt, R. L. (1993) Artifactual variation in<br />

randomly amplified polymorphic DNA banding patterns. <strong>Bio</strong>Techniques 14, 214–217.<br />

7. Meunier, J. R. and Grimont, P. A. (1993). Factors affecting reproducibility of random<br />

amplified polymorphic DNA fingerprinting. Res. Microbiol. 144, 373–379.<br />

8. Burt, A., Carter, D. A., Koenig, G. L., White, T. J., and Taylor, J. W. (1996) Molecular<br />

markers reveal cryptic sex in the human pathogen Coccidioides immitis. Proc. Natl. Acad.<br />

Sci. USA 93, 770–773.<br />

9. Graser, Y., Volovsek, M., Arrington, J., Schonian, G., Presber, W., Mitchell, T. G., et<br />

al. (1996) Molecular markers reveal that population structure of the human pathogen<br />

Candida albicans exhibits both clonality and recombination. Proc. Natl. Acad. Sci. USA<br />

93, 12,473–12,477.<br />

10. Mondon, P., Brenier, M. P., Symoens, F., Rodriguez, E., Coursange, E., Chaib, F., et al.<br />

(1997) Molecular typing of Aspergillus fumigatus strains by sequence-specific DNA primer<br />

(SSDP) analysis. FEMS Immunol. Med. Microbiol. 17, 95–102.<br />

11. Weaver, K. R., Caetano-Anolles, G., Gresshoff, P. M., and Callahan, L. M. (1994) Isolation<br />

and cloning of DNA amplification products from silver-stained polyacrylamide gels.<br />

<strong>Bio</strong>Techniques 16, 226–227.<br />

12. Blanchard, M. M., Taillon-Miller, P., Nowotny, P., and Nowotny, V. (1993) PCR buffer<br />

optimization with uniform temperature regimen to facilitate automation. PCR Methods<br />

Appl. 2, 234–240.<br />

13. Bej, A. K., Mahbubani, M. H., Boyce, M. J., and Atlas, R. M. (1994). Detection of<br />

Salmonella spp. in oysters by PCR. Appl. Environ. Microbiol. 60, 368–373.<br />

14. Bassam, B. J., Caetano-Anolles, G., and Gresshoff, P. M. (1992) DNA amplification<br />

fingerprinting of bacteria. Appl. Microbiol. <strong>Bio</strong>technol. 38, 70–76.<br />

15. Schierwater, B. and Ender, A. (1993) Different thermostable DNA polymerases may<br />

amplify different RAPD products. Nucleic Acids Res. 21, 4647– 4648.<br />

16. Saiki, R. K. (1989) The design and optimization of the PCR, in PCR Technology: Principles<br />

and Applications for DNA Amplification. (Erlich, H. A., ed.), Stockton Press, New York,<br />

pp. 7–16.<br />

17. Yu, K. and Pauls, K. P. (1992) Optimization of the PCR program for RAPD analysis.<br />

Nucleic Acids Res. 20, 2606.<br />

18. Dassanayake, R. S. and Samaranayake, L. P. (2000). Characterization of the genetic<br />

diversity in superficial and systemic human isolates of candida parapsilos by randomly<br />

amplified polymorphic DNA (RAPD). APMIS 108, 153–160.


Microsphere-Based SNP Genotyping 123<br />

24<br />

Microsphere-Based Single Nucleotide<br />

Polymorphism Genotyping<br />

Marie A. Iannone, J. David Taylor, Jingwen Chen, May-Sung Li,<br />

Fei Ye, and Michael P. Weiner<br />

1. Introduction<br />

Single nucleotide polymorphisms (SNPs) are single base differences in genomic<br />

DNA (1). These single-base mutations, estimated to occur every 1000 bases, are<br />

thought to represent the most common form of genetic variation in the human genome<br />

(2). Several million SNPs have been identified (3). High-throughput analysis of these<br />

variations will be required to understand their contribution to disease. We outline a<br />

method that uses solution-based oligonucleotide ligation assay (OLA) (4) or singlebase<br />

chain extension (SBCE) (5,6) for allele discrimination followed by hybridization<br />

to fluorescently encoded microspheres. Flow cytometric analysis of the microspheres’<br />

fluorescent profile yields rapid and accurate SNP genotyping (7,8).<br />

Allele discrimination by OLA or SBCE uses (1) polymerase chain reaction (PCR)-<br />

amplified genomic DNA that encompasses the SNP to be queried (‘target’ DNA); (2)<br />

a synthetic “capture” oligonucleotide probe; and (3) a fluorescent “reporter” (Fig. 1).<br />

In the allele discrimination reaction, an enzyme is used to covalently couple a reporter<br />

molecule to the capture probe in a target-dependent fashion. Each capture probe contains<br />

a sequence that is complementary to the target sequence and a ZipCode sequence that<br />

will associate the genotype result with a specific microsphere population. In OLA, a<br />

DNA ligase covalently couples the fluorescent reporter (a short oligonucleotide) to the<br />

capture probe if the capture and reporter probes correctly match the target DNA. In<br />

SBCE, a DNA polymerase adds a labeled dideoxynucleotide to the capture probe.<br />

After thermal cycling to amplify the signal on the capture probes, the enzymatically<br />

reacted capture probes are incubated with a suspension of up to 100 populations of<br />

fluorescently encoded microspheres where each population is uniquely identified by<br />

its fluorescent profile (9–11). Each microsphere population is covalently coupled to a<br />

different complementary ZipCode (cZipCode) oligonucleotide sequence that associates<br />

it with a capture probe and specific SNP allele. Flow cytometric analysis of the<br />

microspheres simultaneously identifies both the microsphere type and the fluorescent<br />

signal associated with the SNP genotype.<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

123


124 Iannone et al.<br />

Fig. 1. Diagram of microsphere-based SNP genotyping by OLA or SBCE. Allele discrimination<br />

by OLA or SBCE uses PCR-amplified genomic DNA that encompasses the SNP (target<br />

DNA), synthetic capture oligonucleotide probe (capture probes extend up to the polymorphic<br />

base for OLA and end prior to the polymorphic base for SBCE), and a fluorescent reporter.<br />

Each capture probe contains both a sequence that is complementary to the target sequence and a<br />

unique 25-base ZipCode sequence that will link the genotyping result to a specific microsphere<br />

population. For OLA, the reporter is a short target-complementary oligonucleotide sequence<br />

that ends with a fluorescein molecule. If there is base pairing between the reporter and capture<br />

probe, DNA ligase will covalently couple the fluorescent reporter to the capture molecule. For<br />

the SBCE example shown above, the reporter is a fluorochrome or biotin-coupled ddTTP (or<br />

ddUTP) or ddCTP (each labeled ddNTP is used in a different tube). The DNA polymerase<br />

extends the capture probe by one base. In each case, the probes are thermally cycled to amplify<br />

the signal on the capture probes. A suspension of cZipCode-coupled microsphere populations<br />

is added and the capture probes are hybridized to the microspheres through their ZipCode tails.<br />

After washing, the fluorescent profiles of the microspheres are analyzed by flow cytometry. The<br />

OLA example shows a multiplexed reaction of a single SNP using two microsphere populations,<br />

one for each allele. More extensive multiplexing of SNPs can be conducted by combining more<br />

probes with complimentary microsphere populations. The SBCE reaction shows two uniplexed<br />

reactions (one for each allele). The SBCE reaction may be multiplexed by combining probes<br />

for different SNPs and assaying all A or G alleles in a single tube.<br />

The three fluorescent colors associated with each microsphere (two to identify the<br />

microsphere population and one for the genotyping result) are determined by flow<br />

cytometry. We describe two different flow cytometric systems for SNP genotyping.<br />

The first uses a standard bench-top cytometer (FACSCalibur, BD <strong>Bio</strong>sciences, San<br />

Jose, CA) with a 488-nm laser excitation source. Sixty-four individual popula-


Microsphere-Based SNP Genotyping 125<br />

tions of microspheres, manufactured by the Luminex Corporation (Austin, TX), are<br />

identified by their orange and red fluorescent profile. Reporter fluorescence is green.<br />

The second system uses a microsphere-dedicated flow cytometer, also manufactured<br />

by the Luminex Corp., called the LX-100. The two-laser system of the LX-100<br />

uses a red laser (635 nm) to identify the microspheres (red and near infrared emission)<br />

and a green laser (532 nm) to excite the reporter fluorochrome (orange emission).<br />

One hundred individual microsphere populations are available for this system. The<br />

system is available with an XY platform for sampling directly from 96-well microtiter<br />

plates.<br />

2. Materials<br />

2.1. SNP Genotyping by OLA with Readout<br />

on a FACSCalibur Flow Cytometer<br />

1. 2-[N-morpholino] ethenesulfonic acid (MES; Sigma, St. Louis, MO).<br />

2. Microspheres, which are polystyrene beads with a carboxylated surface. Each population<br />

of microsphere has a unique profile of red and orange fluorescence (Luminex Corp.,<br />

Austin, TX). Use 2.5 × 10 6 microspheres in 62 µL of 0.1 M MES (each population in<br />

a separate volume).<br />

3. 1-ethyl-3-(3-dimethylaminopropyl) carbodiimide hydrochloride (EDC, Pierce, Rockford,<br />

IL).<br />

4. Water containing 0.1% sodium dodecyl sulphate.<br />

5. Water containing 0.02% Tween-20.<br />

6. Tris [hydroxymethyl] aminomethane hydrochloride (10 mM)/ethylenediamine-tetraacetic<br />

acid (1 mM), pH 8.0 (TE).<br />

7. 3.3× SSC: 0.5 M NaCl, 0.05 M Na Citrate, pH 7.0.<br />

8. Template DNA (human genomic DNA, 10–20 ng, lyophilized).<br />

9. Target probe primers. These should be designed to yield 150- to 500-bp products (see<br />

Table 1).<br />

10. AmpliTaq Gold (Applied <strong>Bio</strong>systems, Foster City, CA), 5 U/µL.<br />

11. 10× PCR buffer I (Applied <strong>Bio</strong>systems, Foster City, CA). 10× buffer: 500 mM KCl, 100 mM<br />

Tris-HCl (pH 8.3), 15 mM MgCl 2 , 0.01% gelatin (w/v).<br />

12. dNTPs (Applied <strong>Bio</strong>systems, Foster City, CA); 10 mM total (2.5 mM of each dNTP).<br />

13. Taq DNA ligase (40 U/µL) and 10× ligase buffer (New England <strong>Bio</strong>Labs Inc., Beverly,<br />

MA). 10× buffer: 200 mM Tris HCl (pH 7.6), 250 mM K acetate, 100 mM Mg acetate,<br />

100 mM dithiothreitol, 10 mM NAD, and 1% Triton X-100.<br />

14. Oligonucleotides (see Table 1 and Notes section): cZipCodes (Oligos etc, Bethel, ME);<br />

reporters (Oligos etc, Bethel, ME, or <strong>Bio</strong>source/Keystone, Camarillo, CA); captures<br />

(<strong>Bio</strong>source Keystone, Camarillo, CA); and fluorescein-labeled luciferase complement<br />

(<strong>Bio</strong>source Keystone, Camarillo, CA).<br />

15. NaCl (5 M).<br />

16. 12 × 75-mm polystyrene test tubes (Becton Dickinson Labware, Franklin Lakes, NJ).<br />

17. FACSCalibur flow cytometer (BD <strong>Bio</strong>sciences, San Jose, CA) equipped with Luminex<br />

Lab MAP hardware and software (Luminex Corp., Austin, TX).<br />

18. FlowMetrix Calibration Microspheres (Luminex Corp., Austin, TX).<br />

19. Quantum Fluorescence Kit for MESF units of FITC calibration particles (Sigma, St.<br />

Louis, MO). Add one drop from each of five separate populations of microspheres (each<br />

population has a different known amount of fluorescein molecules) to 1 mL of phosphatebuffered<br />

saline.<br />

20. QuickCal software (Sigma, St. Louis, MO).


Table 1<br />

Oligonucleotides<br />

Description Size (nt) Modifications Sequence *,†<br />

SNP 7 amplicon 188 None GTCCCAAGCT GCATGATTGC TCTTTCTCCT TCTTCCCTGA GTCTCTCTCC<br />

ATGCCCCTCA TCTCTTCCTT TTGCCCTCGC CTCTTCCATC CAYGTCTTCC<br />

AAGGCCTGAT GCATTCATAA GTTGAAGCCC TCCCCAGATC CCCTTGGAGC<br />

CTCTGCCCTC CTCCAGCCCG GATGGCTCTC CTCCATTT<br />

Forward PCR primer for SNP 7 120 None GTC CCA AGC TGC ATG ATT GC<br />

Reverse PCR primer for SNP 7 120 None AAA TGG AGG AGA GCC ATC CG<br />

cZipCode 14 with 20 nt Luciferase tag 145 5′ amino, C15 spacer CAG GCC AAG TAA CTT CTT CGG GAT TGC ACC GTC AGC ACC<br />

ACC GAG<br />

Complement to Luciferase tag 120 5′ biotin or FITC CGA AGA AGT TAC TTG GCC TG<br />

OLA reporter for SNP 7 114 5′ PO4, 3′ FITC GTC TTC CAA GGC CT<br />

OLA capture probe for C allele with 150 None CTC GGT GGT GCT GAC GGT GCA ATC CTT TTG CCC TCG CCT<br />

ZipCode 14 CTT CCA TCC Ac<br />

OLA capture probe for T allele with 150 None CTC GGT GGT GCT GAC GGT GCA ATC CTT TTG CCC TCG CCT<br />

ZipCode 14 CTT CCA TCC At<br />

SBCE capture probe with ZipCode 14 149 None CTC GGT GGT GCT GAC GGT GCA ATC CTT TTG CCC TCG CCT<br />

CTT CCA TCC A<br />

* All sequences are written 5′ to 3′.<br />

† The polymorphic base at the 3′ end of the OLA capture probe sequence is shown in lower case.<br />

126 Iannone et al.


Microsphere-Based SNP Genotyping 127<br />

2.2. SNP Genotyping by SBCE with Readout<br />

on the LX-100 Flow Cytometer<br />

1. Microspheres (as described in Subheading 2.1.).<br />

2. Template DNA (as described in Subheading 2.1.).<br />

3. Target Probe Primers (as described in Subheading 2.1.).<br />

4. Oligonucleotides (see Table 1 and Notes section): cZipCodes (as described in Subheading<br />

2.1.); captures (as described in Subheading 2.1.); and biotin-labeled luciferase<br />

complement: (Keystone <strong>Bio</strong>source, Camarillo, CA).<br />

5. Shrimp alkaline phosphatase (2 U/µL) (Amersham <strong>Bio</strong>sciences, Piscataway, NJ).<br />

6. Escherichia coli Exonuclease I (10 U/µL) (Amersham <strong>Bio</strong>sciences, Piscataway, NJ).<br />

7. 2× SBCE reaction mix: (Use 10 µL per reaction) 160 mM Tris-HCl (pH 9.0), 4 mM MgCl 2 ,<br />

50 nM of each capture probe, 2.4 units of AmpliTaq FS (Applied <strong>Bio</strong>systems, Foster<br />

City, CA), 2 µM of the allele-specific biotin-labeled ddNTP (PerkinElmer Life Sciences,<br />

Inc., Boston, MA), and 2 µM each of the other three unlabeled ddNTPs (Amersham<br />

<strong>Bio</strong>sciences, Piscataway, NJ).<br />

8. 1× SSC containing 0.02% Tween-20.<br />

9. Streptavidin R-phycoerythrin conjugate, 0.1 mg/mL in phosphate-buffered saline, pH 7.2<br />

(SA-PE, Molecular Probes, Eugene, OR).<br />

10. LX-100 flow cytometer (Luminex Corp., Austin, TX), equipped with an XY plate<br />

sampler.<br />

11. Instrument calibration particles for the LX-100 (CL1/CL2 and Reporter Calibrator<br />

Microspheres; Luminex Corp., Austin TX).<br />

3. Methods<br />

3.1. Coupling of cZipCodes to Microspheres (for Microspheres Analyzed<br />

on Either the FACSCalibur or the LX-100)<br />

1. Combine 50 µL of microspheres (2.5 × 10 6 microspheres in 0.1 M MES) with 1 µL of<br />

amino-modified cZipCode oligonucleotide (1 mM in water).<br />

2. At two separate times, add 10 µL of 30 mg/mL EDC in water to the microsphere mixture<br />

(at the beginning of the incubation and then after 30 min).<br />

3. Incubate for 60 min at room temperature with occasional mixing and sonication to keep<br />

the microspheres unclumped and in suspension.<br />

4. Add 200 µL of water containing 0.1% SDS. Vortex and centrifuge 5 min at 1100g.<br />

Carefully remove the supernatant. Add 200 µL of water containing 0.02% Tween 20.<br />

Vortex and centrifuge at 1100g.<br />

5. Remove the supernatant and resuspend the microspheres in 200 µL of TE and store at<br />

4°C (stable for 6 mo).<br />

3.2. Determination of cZipCode Coupling Efficiency (for Microspheres<br />

Analyzed on Either the FACSCalibur or the LX-100)<br />

1. Combine 10,000 coupled microspheres with 3 pmol of fluorescein-labeled luciferase<br />

complement (for microspheres to be run on a conventional cytometer) or biotin-labeled<br />

luciferase complement (for microspheres to be run on the LX-100) in 0.1 mL of 3.3× SSC.<br />

The different microsphere populations may be multiplexed at this point.<br />

2. Heat the microsphere suspension for 2 min at 96°C to denature any secondary structure.<br />

3. Incubate for 30 min at 45°C.<br />

4. Add 200 µL of 1× SSC containing 0.02% Tween-20. Vortex and centrifuge 3 min at<br />

1100g. Carefully remove the supernatant and resuspend in 300 µL of 1× SSC containing<br />

0.02% Tween-20.


128 Iannone et al.<br />

5. For microspheres to be analyzed on the LX-100, incubate an additional 30 min with 5 µL<br />

of SA-PE reagent at room temperature in the dark. Analyze orange fluorescence associated<br />

with the microspheres on the LX-100 without washing. Effective coupling reactions<br />

analyzed on our LX-100 yield 2000 to 4000 mean fluorescent intensity (MFI) units.<br />

6. For analysis of green fluorescence associated with the microsphere populations, analyze on<br />

the FACSCalibur flow cytometer and convert MFI values to molecules equivalent soluble<br />

fluorochrome of fluorescein (MESF) (see Subheading 3.7.). Effective coupling reactions<br />

yield >100,000 MESF after background fluorescence contributed by the microspheres<br />

alone has been subtracted. The corrected MESF value will determine the number of<br />

molecules of cZipCode coupled per microsphere.<br />

3.3. Generation of Target Probes by PCR Amplification<br />

(for Either OLA or SBCE)<br />

1. In a Polyfiltronics 96-well plate, amplify 10 to 20 ng of genomic DNA per well in 15- to<br />

30-µL reaction volumes. Each PCR reaction should contain 1.5 units of AmpliTaq Gold,<br />

400 µM dNTPs, 200 µM forward PCR primer, and 200 µM reverse PCR primer in 1× PCR<br />

buffer I (Applied <strong>Bio</strong>systems, Foster City, CA). PCR amplifications are uniplexed.<br />

2. Set the thermal cycler to heat 10 min at 95°C min to activate the DNA polymerase,<br />

followed by 40 three-temperature amplification cycles holding at 94, 60, and 72°C for 30 s<br />

each and ending with an additional 5-min extension at 72°C.<br />

3. Hold samples at 4°C following completion of the reaction.<br />

3.4. OLA<br />

1. Incubate the following in a total volume of 10 µL: 1× ligase buffer, 0.1 pmol of each<br />

capture oligonucleotide, 5 pmol of each reporter oligonucleotide, 3 to 20 ng of each<br />

dsDNA target probe (as determined by Picogreen staining, Molecular Probes, Eugene,<br />

OR), and 10 U Taq DNA ligase. Ligation reactions are multiplexed.<br />

2. Incubate in a thermal cycler by heating to 96°C for 2 min, followed by 30 cycles of a<br />

two-step reaction (denaturation at 94°C for 15 s followed by ligation at 37°C for 1 min).<br />

3. Hold samples at 4°C when the cycles are complete.<br />

3.5. Hybridization of Capture Probes to Microspheres after OLA<br />

1. Add cZipCode-coupled microsphere populations (5000 microspheres of each population)<br />

to the ligation reaction (microsphere populations are combined at this step).<br />

2. Adjust the salt concentration to 500 mM NaCl by adding a small volume of 5 M NaCl.<br />

3. Heat the mixture to 96°C for 2 min in a thermal cycler and incubate at 45°C from 2 h<br />

to overnight.<br />

4. Wash microspheres with 200 µL of 1× SSC containing 0.02% Tween-20 by centrifuging<br />

at 1100g for 5 min.<br />

5. Resuspend the microsphere suspensions in 300 µL of 1× SSC containing 0.02% Tween-20<br />

just before flow cytometric analysis.<br />

3.6. Flow Cytometric Analysis of Microspheres Hybridized<br />

to OLA Products on the FACSCalibur<br />

1. Transfer the microsphere suspensions to 12 × 75-mm polystyrene test tubes.<br />

2. Optimize the settings on the flow cytometer to analyze the microspheres using FlowMetrix <br />

Calibration Microspheres in conjunction with Luminex software.<br />

3. Acquire a minimum of 100 microspheres from each population per tube.<br />

4. Analyze a separate tube of calibration particles (Quantum Fluorescence Kit for MESF<br />

units of FITC) using the same instrument settings.


Microsphere-Based SNP Genotyping 129<br />

3.7. Analysis of FACSCalibur Data<br />

1. Convert all green fluorescence measurements from MFI to MESF by either manually<br />

creating a calibration curve from the fluorescent intensities of the calibration particles<br />

or by using QuickCal software.<br />

2. Adjust raw MESF values by subtracting the microsphere alone control MESF values<br />

to eliminate microsphere-contributed background fluorescence. This is necessary, because<br />

each microsphere population in the 64-microsphere set emits a small amount of fluorescence<br />

in the green reporter channel. Microsphere corrections range from 1000 to<br />

30,000 MESF, depending on the microsphere population.<br />

3. To adjust for tube-to-tube variability in the wash step, one may include a microsphere<br />

population with no cZipCode attached. Subtract the adjusted MESF (as described previously)<br />

of this negative control microsphere from the MESF of every microsphere type in<br />

that particular tube to normalize the data.<br />

4. Merge the data from the two corresponding alleles and graph the results as x-y coordinates.<br />

3.8. SBCE<br />

1. PCR amplification of target probes is performed as described in Subheading 3.3.<br />

2. To degrade the excess PCR primers and dNTPs before the SBCE assay, add 1 unit of<br />

shrimp alkaline phosphatase and 2 units of E. coli Exonuclease I directly to 10 µL<br />

of pooled PCR products (10 to 20 ng of each amplicon) and mix thoroughly.<br />

3. Incubate at 37°C for 30 min and then for 15 min at 80°C to inactivate the enzymes.<br />

4. For each SNP, set up similar reactions differing only by the choice of labeled ddNTP.<br />

Add 10 µL of pooled, treated PCR products to 10 µL of SBCE reaction mix (160 mM<br />

Tris-HCl, pH 9.0; 4 mM MgCl 2 ; 50 nM of each capture probe; 2.4 units of AmpliTaq<br />

FS; 2 µM of the allele-specific biotin-labeled ddNTP; and 2 µM each of the other three<br />

unlabeled ddNTPs).<br />

5. Denature the reactions at 96°C for 2 min and follow with 30 amplification cycles at 94°C<br />

for 30 s, 55°C for 30 s, and 72°C for 30 s.<br />

6. Hold reactions at 4°C.<br />

3.9. Hybridization of Capture Probes to Microspheres after SBCE<br />

1. To the wells of a standard 96-well microtiter plate in a total volume of 30 µL, add 1000<br />

of each cZipCode-coupled microsphere population to 20 µL of SBCE reaction mixture.<br />

Adjust salt concentration to 500 mM NaCl and 13 mM EDTA.<br />

2. Incubate the mixture at 40°C for 1 h.<br />

3. Wash the microspheres with 150 µL of 1× SSC containing 0.02% Tween 20.<br />

4. Centrifuge for 5 min at 1100g and remove the supernatants.<br />

5. Resuspend microspheres in 60 µL of 1× SSC containing 0.02% Tween 20.<br />

6. Add 5 µL of SA-PE to the microsphere-hybridized SBCE reaction products.<br />

7. Incubate the mixture for 30 min at room temperature.<br />

3.10. Flow Cytometric Analysis on the LX-100 and Data Analysis<br />

1. Optimize the settings on the LX-100 to analyze the microspheres using Luminex calibration<br />

particles in conjunction with Luminex software.<br />

2. For each microtiter well, analyze a minimum of 30 microspheres of each population.<br />

3. Adjust the raw MFI values for microsphere background fluorescence by subtracting microsphere<br />

alone control MFI values from the MFI value of each corresponding microsphere sample.<br />

This is necessary because each microsphere population in the 100-microsphere set emits<br />

a small amount of fluorescence into the orange reporter channel. Microsphere corrections<br />

range from 1 to 100 MFI, depending upon the microsphere population.


130 Iannone et al.<br />

4. To adjust for well-to-well variability in the wash step, each well may contain a microsphere<br />

population with no cZipCode attached. Subtract the adjusted MFI (as described above)<br />

of this negative control microsphere from the MFI of every microsphere type in that<br />

particular well.<br />

5. Merge the data from the two corresponding alleles and graph the results as x-y coordinates.<br />

4. Notes<br />

1. Oligonucleotide probes for OLA: (1) cZipCodes. We have designed our 58 different<br />

cZipCode sequences to include: (a) 5′ amine group, an 18-atom spacer (CH 3 CH 2 O) 6<br />

to minimize any potential interactions between the oligonucleotide sequence and the<br />

microsphere surface; (b) a common 20-base sequence from luciferase cDNA (5′-CAG<br />

GCC AAG TAA CTT CTT CG-3′) to test for oligonucleotide coupling efficiency; and<br />

(c) a 25-base, non-crossreacting cZipCode sequence derived from the Mycobacterium<br />

tuberculosis genome (7,8) to link each allele of SNP to a particular microsphere population.<br />

The cZipCode sequences have GC-contents between 56 and 72% and predicted Tm values<br />

of 61 to 68°C. Although this chapter does not outline the methodology, biotinylated<br />

cZipCodes may be coupled to Lumavidin-coated microspheres. (2) Reporters. These<br />

oligonucleotides are designed to hybridize to the target sequence immediately downstream<br />

of the capture probe. They are generally 8 to 20 bases in length and contain a 5′ phosphate<br />

group and a 3′ fluorescein modification. The phosphate group is required as a substrate<br />

for the ligase enzyme. Reporter probe Tm range from 36 to 40°C. (3) Captures. The 5′<br />

end of each capture probe contains a 25-nucleotide ZipCode sequence and the 3′ end<br />

contains a 20 to 25 base target-specific sequence that extends to the polymorphic base.<br />

This orientation insures optimal ligase fidelity (12,13). Because each SNP has a minimum<br />

of two polymorphic bases, each SNP will require at least two different OLA capture<br />

probes. If the alleles are assayed in the same reaction volume, these capture probes will<br />

also require different ZipCode sequences. Target-specific capture sequences have a Tm<br />

of 51 to 56°C. (4) Fluorescein-labeled luciferase complement. A 5′ fluoresceinated oligo<br />

complementary to the 20 nucleotides of luciferase sequence in each cZipCode.<br />

2. We have found that only 58 of the 64 red–orange populations Luminex provides could be<br />

run on our FACSCalibur with the Luminex software. The number of useable populations<br />

of microspheres may vary depending upon the cytometer model being used. We have used<br />

all 100 microsphere populations on the LX-100.<br />

3. We have successfully multiplexed over 50 SBCE reactions.<br />

4. Oligonucleotide probes for SBCE: (1) cZipCodes (same as OLA); (2) Captures. Similar to<br />

OLA captures with the exception that target-specific component of SBCE capture probes<br />

is designed to stop just short of the polymorphic base. For SBCE, each SNP will require<br />

only one capture probe. Different alleles are assayed in different reaction volumes. Each<br />

volume will have a different biotin-labeled ddNTP plus the other three unlabeled ddNTPs.<br />

(3) <strong>Bio</strong>tin-labeled luciferase complement. A 5′ biotin-modified oligo complementary to<br />

the 20 nucleotides of luciferase sequence found in each cZipCode.<br />

5. Figure 2 shows representative SNP genotyping results from approx 100 patients for 6 A/G<br />

SNPs analyzed by SBCE on the LX-100. Homozygous and heterozygous clusters are<br />

readily discernible.<br />

Fig. 2. (see facing page) SNP genotyping by SBCE with analysis on the LX-100. SNP<br />

genotyping results from approx 96 patients for 6 A/G SNPs. <strong>Bio</strong>tinylated ddNTPs were used<br />

(one plate for G alleles, one for A alleles) followed by incubation with SA-PE. The results shown<br />

come from an experiment where 35 SNPs were genotyped in two microtiter plates. The points in<br />

the lower left corner represent either negative controls or failed PCR reactions.


Microsphere-Based SNP Genotyping 131<br />

131


132 Iannone et al.<br />

6. Target concentration has a direct impact on signal intensity. Each panel in Fig. 2 includes<br />

one negative control point. Additional points shown in the lower left are most likely the<br />

result of PCR failure.<br />

7. High salt concentrations help hybridization efficiency. However, we have found that<br />

when analyzing samples on the LX-100, salt concentrations ≥400 mM result in optical<br />

disturbances. This may be caused by the different refractive indices of the high ionic<br />

strength core stream and the low ionic strength sheath fluid (14).<br />

8. We have reserved one microsphere population and its associated cZipCode for use<br />

with an SBCE positive control. Each SBCE reaction includes a short (40mer) synthetic<br />

oligonucleotide target with a 4-fold degenerate position near its center and a complementary<br />

capture probe to insure incorporation of reporter signal for any nucleotide assayed. Use<br />

of this positive control in every SBCE assay well provides an excellent internal standard<br />

to assess well-to-well reaction variability and can be used to normalize signal intensities<br />

generated across a plate of 96 samples (14).<br />

9. We have also used Rhodamine-6G (R6G)-labeled ddNTPs with the SBCE system on the<br />

LX-100. Although using this directly coupled fluorochrome saves an additional wash and<br />

incubation step, the fluorescent intensities were not as bright as PE. This is not unexpected<br />

since the quantum efficiency of PE is much better than R6G.<br />

10. Genotyping by OLA permits allele multiplexing and has no requirement for shrimp<br />

alkaline phosphatase and Exonuclease I pretreatment of PCR target probes. The advantage<br />

of SBCE is that only one capture probe is required for each allele, thus saving on costs.<br />

The accuracy of genotyping using OLA or SBCE is >99% (7,8).<br />

11. The inclusion of the detergent Tween 20 in the buffer helps reduce the loss of microspheres<br />

during wash steps.<br />

12. Because the LX-100 uses a dual-laser system for microsphere identification and reporter measurement,<br />

spectral compensation is not required on the LX-100 as on the FACSCalibur.<br />

13. Our current automated genotyping facility uses SBCE with 52 microsphere populations.<br />

Enhanced throughput may be realized by incorporating multiplexed PCR amplifications.<br />

14. Conversion of MFI to MESF, although not required, offers several advantages. These<br />

advantages include (1) the use of a standard fluorescence unit; (2) the ability to compare<br />

data between experiments and instruments; and (3) normalization of signal variability in an<br />

instrument over time (caused by laser power shifts or PMT decline). Although not outlined<br />

in this chapter, data from the LX-100 may also be converted to MESF using calibration<br />

microspheres for PE. We have used QuantiBRITE PE beads from BD <strong>Bio</strong>sciences for<br />

this purpose.<br />

15. Although this chapter focuses on the use of Luminex microspheres for multiplexed analysis,<br />

there are other products available for developing multiplexed assays using a conventional<br />

bench-top flow cytometer such as the FACSCalibur (see http://www.spherotech.com,<br />

http://www.bangslabs.com). These products use one fluorescent parameter for the identification<br />

of the microsphere populations and have a more limited level of multiplexing. For<br />

manual acquisition of multiplexed assays on a conventional bench-top cytometer, (1) make<br />

sure microspheres can be seen in forward versus side scatter and gate to exclude doublets;<br />

(2) set the PMTs of fluorescent parameters to log acquisition; (3) adjust the PMT settings<br />

of fluorescent parameters so that unstained and brightly stained microspheres are not offscale;<br />

and (4) adjust compensation settings to subtract reporter fluorescence from the other<br />

fluorescent channels. When properly compensated, microsphere populations with bright<br />

reporter staining will not move out of the regions that identify their population.<br />

16. Microsphere-based SNP genotyping by OLA or SBCE is an accurate and rapid method<br />

for the analysis of SNPs.


Microsphere-Based SNP Genotyping 133<br />

References<br />

1. Landegren, U., Nilsson, M., and Kwok, P. Y. (1998) Reading bits of genetic <strong>info</strong>rmation:<br />

Methods for single-nucleotide polymorphism analysis. Genome Res. 8, 769–776.<br />

2. Cooper, D. N., Smith, B. A., Cooke, H. J., Niemann, S., and Schmidtke, J. (1985) An<br />

estimate of unique DNA sequence heterozygosity in the human genome. Hum. Genet.<br />

69, 201–205.<br />

3. Marshall, E. (1999) Drug firms to create public database of genetic mutations. Science<br />

284, 406– 407.<br />

4. Abravaya, K., Carrino, J. J., Muldoon, S., and Lee, H. H. (1995) Detection of point<br />

mutations with a modified ligase chain reaction (Gap-LCR). Nucleic Acids Res. 23,<br />

675–682.<br />

5. Syvanen, A. C. (1999) From gels to chips: “Minisequencing” primer extension for analysis<br />

of point mutations and single nucleotide polymorphisms. Hum. Mutat. 13, 1–10.<br />

6. Syvanen, A. C., Aalto-Setala, K., Harju. L., Kontula, K., and Soderlund, H. (1990) A<br />

primer-guided nucleotide incorporation assay in the genotyping of apolipoprotein E.<br />

Genomics 8, 684–692.<br />

7. Iannone, M. A., Taylor, J. D., Chen, J., Li, M.-S., Rivers, P., Slentz-Kesler, K. A., et al.<br />

(2000) Multiplexed single nucleotide polymorphism genotyping by oligonucleotide ligation<br />

and flow cytometry. Cytometry 39, 131–140.<br />

8. Chen, J., Iannone, M. A., Li, M-S., Taylor, J. D., Rivers, P., Nelson, A. J., et al. (2000) A<br />

micropshere-based assay for multiplexed single nucleotide polymorphism analysis using<br />

single base chain extension. Genome Res. 10, 549–557.<br />

9. Fulton, R. J., McDade, R. L., Smith, P. L., Kienker, J. J., and Kettman, J. R. (1997) Advanced<br />

multiplexed analysis with the FlowMetrix system. Clin. Chem. 43, 1749–1756.<br />

10. Kettman, J. R., Davies, T., Chandler, D., Oliver, K. G., and Fulton, R. J. (1998) Classification<br />

and properties of 64 multiplexed microsphere sets. Cytometry 33, 234–243.<br />

11. McDade, R. L. and Fulton, R. L. (1997) True multiplexed analysis by computer-enhanced<br />

flow cytometry. Med. Dev. Diag. Indust. 19, 75–82.<br />

12. Husain, I., Tomkinson, A. E., Burkhart, W. A., Moyer, M. B., Ramos, W., Mackey, Z. B., et<br />

al. (1995) Purification and characterization of DNA ligase III from bovine testes. Homology<br />

with DNA ligase II and vaccinia DNA ligase. J. <strong>Bio</strong>l. Chem. 270, 9683–9690.<br />

13. Luo, J., Bergstrom, D. E., and Barany, F. (1996) Improving the fidelity of Thermus<br />

thermophilus DNA ligase. Nucleic Acids Res. 24, 3071–3078.<br />

14. Taylor, J. D., Briley, D., Nguyen, Q., Long K., Iannine, M. A., Li, M. S., et al. (2001)<br />

Flow cytometric platform for high-throughput single nucleotide polymorphism analysis.<br />

<strong>Bio</strong>echniques 30, 661–669.


134 Iannone et al.


Ligase Chain Reaction 135<br />

25<br />

Ligase Chain Reaction<br />

William H. Benjamin, Jr., Kim R. Smith, and Ken B. Waites<br />

1. Introduction<br />

The ligase chain reaction (LCR) is one of many techniques developed in recent years<br />

to detect specific nucleic acid sequences by amplification of nucleic acid targets. The<br />

LCR has been used for genotyping studies to detect tumors and identify the presence<br />

of specific genetic disorders such as sickle cell disease caused by known nucleotide<br />

changes that occur as a result of point mutations and has now become widely used<br />

in infectious disease detection, both in the diagnostic and research settings, primarily<br />

focusing on infections caused by microbes that have proven difficult to detect by<br />

traditional culture techniques. The LCR is now recognized as the method of choice<br />

for detection of urogenital infections due to Chlamydia trachomatis because of its<br />

greater sensitivity as compared to traditional cell culture or nonamplified DNA probes<br />

or antigen-detection assays. Other uses of the LCR have also been reported (1–8). When<br />

used for detection of infectious diseases, amplification tests such as the LCR have the<br />

additional advantages in that they do not require viable organisms in a specimen, a<br />

single specimen can be used to detect multiple different pathogens, provided suitable<br />

primers are available, and easily obtained specimens such as urine can be used for<br />

diagnostic purposes, making screening of large numbers of persons practical, as well as<br />

facilitating research to better understand the epidemiology of specific diseases.<br />

Ligation of adjacent oligonucleotides while hybridized to a template was first<br />

investigated by Besmer et al. in 1972 (9). The first use of ligases to join oligonucleotides<br />

as a means to differentiate sequence variants was reported in 1988 (10,11) and the first<br />

“real” LCR utilizing cyclical denaturation hybridization and ligation of two pairs of<br />

oligomers was described in 1989 (12). Thermostable ligases that eliminate the necessity<br />

of adding ligase after each denaturation step were introduced by Barany in 1991 (1).<br />

Simple LCR consists of two complementary oligonucleotide pairs (four oligonucleotides<br />

20–35 nucleotides each) that are homologous to adjacent sequences on the<br />

target DNA, as opposed to two used in the polymerase chain reaction (PCR) assay.<br />

The adjacent pairs are ligated when they hybridize to the complementary sequence<br />

next to each other in a 3′ to 5′ orientation on the same strand of the target DNA.<br />

The 5′ nucleotide of the ends of the primers to be ligated must be phosphorylated.<br />

Newly ligated oligonucleotides become targets in subsequent cycles so logarithmic<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

135


136 Benjamin, Smith, and Waites<br />

amplification occurs. The two complementary pairs can also, at low frequency, be blunt<br />

end ligated to each other and serve as template for amplification even though no target<br />

sequence was present in the original sample. Early work showed at least 10-fold greater<br />

efficiency of ligation of perfectly matched compared to a mismatched nucleotide at<br />

the ligation junction (1,8,12).<br />

LCR-based systems have some advantages over the PCR-based amplification systems.<br />

Because there is no newly synthesized DNA, misincorporated nucleotides are not<br />

replicated in the product allowing amplification of a different sequence than that found<br />

in the target nucleic acid. The LCR reactions are also more specific for the 3′ nucleotide<br />

allowing for higher discriminatory power against mismatches at a single chosen site (1,3).<br />

Thus, LCR is very useful for determining the nucleotide at a specific site such as a single<br />

base change, e.g., single-nucleotide polymorphisms (SNPs) used in mapping complex<br />

genomes. The LCR cycle has only two short steps allowing for shorter amplification<br />

times. The usually small target of LCR, 36 to 60 nucleotides, does not require highquality<br />

large fragment nucleic acids (13). The commercial LCR kit, the Abbott LCx<br />

System (Abbott Diagnostics, Abbott Park, IL, USA) is less affected by inhibitors in some<br />

specimens, such as fresh urines compared to PCR (14). However, the LCR is still subject<br />

to contamination that is at least as much of a problem as for the PCR.<br />

There are several modifications to the simple LCR that have been used to overcome<br />

some of the problems that are inherent in this method. Owing to the specificity of the<br />

ligase reaction for perfect matches that is complicated by the blunt end ligation of<br />

double stranded probes PCR, the Ligase Detection Reaction (LDR) was developed<br />

(15). The amplification step for LDR is PCR using outside primers, then only one<br />

pair of adjacent primers is used to detect the proper sequence in the PCR product<br />

(1,7). Because there are no double stranded oligonucleotides to blunt end ligate, the<br />

background is very low and the linear detection of the PCR amplified sequence can<br />

be detected using ligation of oligonucleotides. Gap LCR utilizes a few base pair gaps<br />

between the two pair of oliogonucleotides on the primer, thus requiring a thermostable<br />

polymerase (Taq polymerase) to fill in the gap before ligation of adjacent primers can<br />

occur. The Gap LCR prevents amplification of the blunt end ligated primers because the<br />

oligonucleotide pairs cannot be amplified on blunt end ligated products (16,17).<br />

1.1. Ligase<br />

There are four thermostable ligases that can be used in LCR reactions. Taq ligase<br />

was the first described and is commercially available from New England <strong>Bio</strong>labs<br />

(www.uk.neb.com/neb/index.html) (Beverly, MA). Stratagene (www.stratagene.<br />

com) (La Jolla, CA) produces Pfu which is advertised as having higher ligation<br />

specificity and lower background than Tth that is marketed by Abbott Diagnostics<br />

(www.abbottdiagnostics.com) (Abbott Park, IL) and is used in their LCx diagnostic kit.<br />

Amplicase is sold by Epicenter Technologies (www.epicentre.com) (Madison, WI).<br />

1.2. Probe Design<br />

The most successful probes are 18 to 30 bases in length. Probes with high GC<br />

content or those that form stable secondary structures or dimers should be avoided<br />

if possible. It is difficult to predict activity, so empiric testing is still necessary (18).<br />

The 5′ nucleotide of the adjacent end must be phosphorylated to promote ligation.


Ligase Chain Reaction 137<br />

The probe candidates must be tested for sensitivity by determining the lower level of<br />

detection of samples containing organisms of interest or their DNA. Specificity must<br />

be determined by testing related organisms, as well as other organisms likely to be<br />

found in specimens to be tested.<br />

1.3. Target Sequence Selection<br />

The sequence should be unique to the organism to be detected and multiple copies<br />

per organism increases sensitivity. If the ligase reaction is to be used to discriminate<br />

between single base pairs, the ultimate 3′ base is most sensitive, whereas the 5′ base<br />

is also relatively sensitive. Allele-specific PCRs used to detect single nucleotide<br />

polymorphisms are often not discriminating enough to differentiate between SNPs,<br />

whereas the ligase reactions are more discriminating against mismatches, especially<br />

on either side of the ligation site (1). All of the possible mismatches are discriminated<br />

against, but G-T and T-T less well, 1.5% as efficient as the matching nucleotide, while<br />

other mismatches were 80%. When the mismatch was<br />

on the 5′ side the previous two and A-C, A-A, C-A, G-A and T-T, all were all less<br />

discriminated against than when on the 3′ side, same patterns as were found with T4<br />

ligase. Intentionally introducing a mismatch in the third site from the 3′ end of the<br />

probes increased the discriminating power. Nucleotide analogs in the probes in the<br />

2 and 3 location from the 3′ end also increased discriminatory power. Site-directed<br />

mutagenesis was used on Tth and mutants that increased discriminating power 4- and<br />

11-fold were found by Luo et al. (3).<br />

1.4. Detection of Amplification Products<br />

Multiplex LCR using a mixture of probes differing by the 3′ nucleotide involved<br />

in the ligation, that are labeled by being one or two extra bases on the nonligating<br />

5′ end, allows polyacrylamide gel electrophoresis (PAGE) differentiation of the one<br />

or two base changes by differences in migration in PAGE (1). The probes can also be<br />

labeled with several different, easily detected labels such as: 32P, fluorescent labels,<br />

immunologically detectable haptens (digoxigenen), (Roche Molecular Diagnostics,<br />

Indianapolis, IN) and after amplification and electrophoresis the signal is detected<br />

by Southern blot to determine if ligation has taken place. Ligated products can also<br />

be detected by having immuno-capture of one of the probe ends and after washing,<br />

detecting the second probe with an enzyme conjugate. Only ligated product will be<br />

captured and also have the end with the ligand for which the enzyme conjugated<br />

antibody is specific, IMx (Abbott) utilizes this method. The latter method is also<br />

employed in the commercially sold Abbott LCx for C. trachomatis and N. gonorrhoeae.<br />

The comparison of eight different nonradioactive methods of detecting the LCR<br />

products was reported by Winn-Deen (19).<br />

1.5. Contamination Control<br />

One problem with LCR is that the target is amplified, resulting in a contamination<br />

risk. The method commonly used to inactivate PCR products does not work because of


138 Benjamin, Smith, and Waites<br />

the small size of the amplicon in LCR. The potential for contamination requires strict<br />

adherence to physical separation of setup and detection areas and other containment<br />

methods such as bleach treatment of lab benches, Ultraviolet irradiation of setup areas,<br />

unit premixes for setup, as well as use of aerosol barrier pipet tips. The commercial<br />

LCx instrument injects a binary inactivating agent that is capable of inactivating small<br />

amplicons by a factor of up to 109.<br />

1.6. Inhibitors<br />

LCR is less prone to problems with inhibitors from urogenital specimens than PCR<br />

for detection of Chlamydia (14,20). Freeze-thawing and dilution decreased the falsenegative<br />

rate of PCR (21). Possible inhibitors of LCR were removed by a mildly acid<br />

wash to remove CaHPO4 from concentrated specimens being tested for the presence of<br />

acid fast bacteria (22). Potential effect of inhibitors on amplification results mandates<br />

rigorous quality control for all steps of the procedure as described in the technical<br />

procedures presented here.<br />

1.7. LCR Applications<br />

1.7.1. Noncommercialized Methods<br />

Over the past few years, numerous publications have appeared that describe a<br />

number of potential applications of the LCR procedure, each with its own particular<br />

modifications of the basic procedure, using the varied available ligases, probe designs,<br />

relative positions, and detection methods. Table 1 shows references and lists the<br />

methods used for several diverse applications. A basic LCR procedure that can be<br />

adapted to detect nucleic acid of a variety of etiologic agents as well as eukaryotic<br />

polymorphisms, especially SNPs, is presented in further detail in Subheading 2.<br />

1.7.2. Commercially Available LCR Kits*<br />

As an artifact of the way the companies interested in molecular diagnostics have<br />

carved up the field, mostly driven by the ownership of rights to specific procedures,<br />

in the United States, LCR is primarily used to diagnose the sexually transmitted<br />

diseases caused by C. trachomatis and N. gonorrhoeae. Abbott Diagnostics markets<br />

a commercial LCR kit, the LCx, in which four oligonucleotide probes target and<br />

hybridize with a specific complementary single-stranded nucleotide sequence within<br />

the multicopy cryptic plasmid gene present in all serovars of C. trachomatis that is<br />

exposed during sample preparation in which the heating process causes the release<br />

of single-stranded DNA, leaving a gap of a few nucleotides between the probes (23).<br />

Polymerase then fills the gap with nucleotides in the LCR reaction mixture. Once<br />

the gap is filled, thermostable ligase covalently joins the pair of probes to form an<br />

amplification product that is complementary to the original target sequence that then<br />

serves as an additional target sequence for further rounds of amplification. Amplification<br />

occurs when the LCR reaction mixture and sample are incubated in a thermal<br />

cycler. During thermal cycling, the temperature is raised above the melting point of<br />

*Abbott Diagnostics, which marketed the LCx Uriprobe that was FDA approved in 1994, is to be<br />

discontinued in June 2003 and after that time no commercial kits will use the LCR technology.


Ligase Chain Reaction 139<br />

the hybridized amplification product causing it to dissociate from the original target<br />

sequence. Lowering the temperature allows more of the oligonucleotide probes to<br />

hybridize to the targets now available and to be ligated. The temperature continues to<br />

be cycled until sufficient numbers of target amplification product have accumulated.<br />

Ligated product is captured by antibody immobilized onto the surface of microparticles<br />

using a ligand attached to the end of one primer and then detected by an enzymeconjugated<br />

antibody directed at a second reporter molecule at the distal end of the<br />

other primer. Only ligated product with both haptens covalently attached will generate<br />

a chemiluminescent signal. The amplification product accumulates exponentially and<br />

is detected by chemiluminescence on the automated Abbott LCx or IMx Analyzers.<br />

Specimens that can be used include cervical swabs and urethral swabs as well as urine<br />

from either men or women (24–26). The LCx kit also has the advantage of being<br />

multiplexed for the detection of N. gonorrhoeae in the same genitourinary specimen.<br />

The LCx target in N. gonorrhoeae is a 48 base pair DNA sequence in the multicopy Opa<br />

gene that is conserved in all strains studied to date and is specific to N. gonorrhoeae.<br />

There are mixed results on the sensitivity of LCx compared to other methods of<br />

amplification for detection of urogenital infections. One report found PCR more<br />

sensitive than LCx. (27), whereas a second study found LCx more sensitive (28). Stary<br />

et al. compared the sensitivity and specificity of LCx and the Transcription Mediated<br />

Amplification (TMA) assay (GenProbe, Inc., San Diego, CA) using several different<br />

urogenital specimens. They found comparable results with endocervical and vulval<br />

swab samples, male urethral swabs and urine, but found a lower sensitivity with the<br />

TMA method for female first void urine (29). Carroll et al. found the overall sensitivity<br />

of LCx was better than GenProbe PACE 2 for chlamydia with very good specificity<br />

for both (30).<br />

Abbott Diagnostics also makes a Mycobacterium (M.) tuberculosis LCx kit that is<br />

approved for diagnostic use in some European countries but not in the United States<br />

at present. This product is a Gap LCR assay with a few nucleotides that must be filled<br />

before ligation. It targets the protein antigen b genes of M. tuberculosis. This LCx<br />

kit has been evaluated in comparison to culture and/or other amplification methods<br />

such as the Roche PCR and the TMA with generally favorable results. Overall, most<br />

evaluations have found the LCx to perform better for detection of M. tuberculosis in<br />

smear-positive as opposed to smear-negative respiratory specimens (31–34). However,<br />

one investigation (35) found similar high sensitivity and specificity values for the LCx,<br />

exceeding 90% for smear-positive as well as smear-negative specimens, suggesting<br />

that the LCx assay has potential utility as a screening test for the rapid diagnosis of<br />

tuberculosis in high-risk patients.<br />

2. Materials<br />

2.1. Equipment<br />

1. Microcentrifuge capable of speeds of ≥9000g.<br />

2. Vortex mixer.<br />

3. 20°C freezer for sample storage if not processed immediately.<br />

4. LCx Analyzer or other detection system with appropriately labeled reagents specific for<br />

the system employed.<br />

5. LCx Thermocycler.<br />

6. LCx dry bath capable of heating from 60° to 100°C.


140 Benjamin, Smith, and Waites<br />

2.2. Supplies and Reagents<br />

1. LCx kit (if commercial assay is being used, otherwise individual components as listed<br />

in Table 1 and Steps 2–9).<br />

2. Taq ligase or other type as described in Subheading 1.1. with appropriate buffers from<br />

same supplier.<br />

3. Custom probes with primers chosen according to criteria outlined in Subheadings 1.2.<br />

and 1.3.<br />

4. Proteinase K (final concentration 200 µg/ml).<br />

5. Mineral oil.<br />

6. Specimen or analyte.<br />

7. Specimen Collection and Transport System (Uriprobe) containing 0.5 ml transport buffer.<br />

8. Sterile, preservative-free plastic screw-top containers for collection and transport of<br />

specimen (if using commercial LCx).<br />

9. 100 µL aerosol barrier pipet tips and pipets.<br />

3. Methods<br />

3.1. Procedural Precautions<br />

1. Work in laboratory using DNA amplification methods should always flow in a one-way<br />

direction beginning in the specimen preparation and processing area (Area 1), then moving<br />

to the amplification and detection area (Area 2). Do not bring any materials or equipment<br />

from Area 2 into Area 1.<br />

2. Surface cleaning using a 1% (v/v) sodium hypochlorite solution followed by 70% (v/v)<br />

ethanol should be performed on bench tops and pipets prior to beginning the LCR Assay.<br />

3. Chlorine solutions may pit equipment and metal. Use sufficient amounts or repeated<br />

applications of 70% ethanol until chlorine residue is no longer visible.<br />

3.2. Prototype LCR Technique for Detection of Neisseria gonorrhoeae<br />

The type of materials required for LCR assays varies greatly according to the ligase,<br />

amplification, and detection systems used, as well as the primers required that must be<br />

specific for the desired target. Table 1 summarizes major types of LCR procedures and<br />

the various materials and equipment needed to perform the assays, including primers,<br />

ligases, buffers, templates, test conditions, targets, and detection systems. A more<br />

detailed description of the Neisseria gonorrhoeae detection LCR that was eventually<br />

modified and incorporated into the Abbott LCx system is described in Subheadings<br />

3.2.1.–3.3.9. These methods were originally developed by Birkenmeyer and Armstrong<br />

and described in 1992 (36). Although this procedure was developed specifically for the<br />

detection of N. gonorrhoeae DNA, the basic principle can be adapted for the detection<br />

of other microbial agents, provided specific oligonucleotide primers are designed.<br />

3.2.1. Primer Selection<br />

1. Primers should be designed to be homologous to conserved sequences in N. gonorrhoeae<br />

that have mismatches with N. meningitidis, the species most homologous to N. gonorrhoeae.<br />

Because of the empiric nature of developing LCR, multiple primer sites should be<br />

developed. The site demonstrating the best sensitivity and specificity should be chosen and<br />

this must be validated as described previously in Subheading 1.2. In the example given,<br />

three sites were chosen corresponding to sequences immediately upstream of several of<br />

the opa gene family (Opa-2 and Opa-3) and a site downstream of several of the pil gene<br />

family (Pilin-2) (36).


Table 1<br />

Examples of Various Applications and Methods for the Ligase Chain Reaction<br />

References Barany (1) Jurinke (2) Khanna (5) Abravaya (38) Day (7) Reyes (8)<br />

Wilson (37) Luo (3)<br />

Zirvi (6)<br />

Ligase Taq 15 U Pfu 4 U Tth 25 fmol Tth 5000 U Tth 10 U Amplicase 1.5 U<br />

Buffers 20 mM/7.6 20 mM/7.5 20 mM/7.6 10 mM/7.5 20 mM/8.3<br />

Tris<br />

EPPS 50 mM/7.8<br />

KCl 100 mM 20 mM 100 mM 20 mM 50 mM 25 mM<br />

MgCl 2 10 mM 10 mM 10 mM 30 mM 10 mM 10 mM<br />

NaCl 400 mM<br />

EDTA<br />

1 mM<br />

NAD+ 10 mM 1 mM 10 µm 0.5 mM<br />

DTT 10 mM 1 mM 10 mM<br />

NP-40 0.10%<br />

Triton X-100 0.01%<br />

Primers 40 fmol 3.3 pmol 500 fmol (38) 30 nm 600 fmol 3.6 nm<br />

Template 1 fmol (1); 0.74 fmol 500 fmol (3); 100 molecules 10–100 ng 250 ng<br />

6–60 µg (37) 100 fmol (5); in 500 ng genomic<br />

50 fmol (6) DNA<br />

Conditions<br />

Denaturation 94°C, 1 min 92°C, 20 sec 94°C, 15 sec 85°C, 30 sec 95°C, 15 sec 91°C, 30 sec<br />

Lgation 64°C, 4 min 60°C, 40 sec 65°C, 4 min 60°C, 30 sec 65°C, 4 min 55°C, 6 min<br />

No. cycles 20–30 25 20 25 5 (LDR) a 27<br />

Target Sickle cell mutation E. coli lacI Human eIF-4E (3); HIV AZT-resistant Human gene Sickle cell<br />

(1); p53, Ha-ras (37) K-ras (5); human mutants; Chlamydia CYP21 mutation<br />

microsatellites (6) cryptic plasmid<br />

Differential Size-labeled primers Size-labeled Fluorescent-labeled Microparticles Size-labeled Arbitrary<br />

detection (1); 32 P label (37) primers; primers; coated with primers; sequence on<br />

matrix-assisted microparticles anti-hapten; sandwich fluorescent primer for<br />

laser coated with immunoassay labeled capture<br />

desorption/ anti-hapten performed primers; detection hybridization;<br />

ionization using the Abbott in Perkin–Elmer biotinylated<br />

time-of- IMx automated GeneScan primer<br />

flight-mass analyzer sequencing system streptavidinspectrometry<br />

alkaline<br />

phosphatase<br />

conjugate<br />

Ligase Chain Reaction 141<br />

a Ligase detection reaction uses ligase chain reaction to detect previously amplified product.


142 Benjamin, Smith, and Waites<br />

2. The primers for this technique are designed to have a gap of 4 to 5 Gs that would be<br />

filled by adding the single deoxynucleotide dGTP and Taq polymerase that would not add<br />

more nucleotides than are needed to fill the gap. The Pilin-2 locus has 6 bp of nontargeted<br />

DNA added to one end that are complementary to the related primer.<br />

4. The ends of the primers are labeled with fluorescein on one pair of homologous primers (one<br />

5′ one 3′) and biotin on the other nonligating ends of the other two primers. The sequences<br />

of the primers are described in the original publication describing this technique (36).<br />

5. The ligating 5′ ends are phosphorylated to facilitate ligation after the gap is filled.<br />

6. All of the modifications of the oligonucleotides can be ordered from numerous companies<br />

that will custom synthesize oligonucleotides for specific purposes and targets.<br />

3.2.2. Clinical Specimens<br />

1. The type of specimen should be chosen according to the site that the organism is most likely<br />

to be detected in association with disease. For N. gonorrhoeae, urethral or endocervical<br />

swabs are collected and placed into 500 µl of specimen buffer containing 50 mM EPPS<br />

buffer [N-(2-hydroxyethyl)piperazine-N′-(3-propanesulfonic acid)] and 5 mM EDTA<br />

(pH 7.8). The tube can be maintained up to 24 to 48 h at room temperature if necessary<br />

for transport to the laboratory. The swabs are vigorously vortexed prior to removal and the<br />

sample is then frozen at -20°C until DNA extraction.<br />

2. Organism lysis to release DNA is accomplished by adding proteinase K to a final<br />

concentration of 200 µg/ml and incubating for 1 h at 60°C.<br />

3. Boiling for 10 min inactivates the proteinase K, further lyses the bacteria and denatures<br />

the DNA.<br />

4. The cell debris is pelleted by centrifuging for 10 min at 4°C at 13,000g.<br />

5. The supernatants can be stored at –20°C until used.<br />

3.2.3. Amplification and Detection<br />

1. Each 50 µl amplification mixture will contain 4 µl of sample, and a final concentration of:<br />

20 mM Tris-HCl (pH 7.6 @25°C) 25 mM potassium acetate, 10 mM magnesium acetate,<br />

10 mM DTT, and 1 mM NAD and 0.1% Triton X-100. This buffer supplied as 10X the<br />

concentrations needed can be purchased commercially (New England <strong>Bio</strong>labs, Beverly,<br />

MA) along with the Taq ligase.<br />

2. Each of the four oligonucleotides within a set should be diluted in advance and the<br />

appropriate volumes of each added to obtain the 830 fmol of each of the four in a set<br />

in each reaction tube.<br />

3. This mixture is overlaid with sterile mineral oil and heated for 3 min in a boiling water<br />

bath to ensure complete denaturation. After cooling, 1 unit of Amplitaq DNA polymerase<br />

(Perkin-Elmer Cetus, Norwalk, CT) and 15 units of Thermus aquaticus (Taq) DNA ligase<br />

(New England <strong>Bio</strong>labs) is added.<br />

4. A positive control consisting of 270 cell equivalents of N. gonorrhoeae DNA in 80 ng/µl<br />

human placental DNA (Sigma Chemicals, Saint Louis, MO) and a negative control must<br />

be included in each run.<br />

5. Gap LCR is performed in a thermocycler. The prototype technique for N. gonorrhoeae<br />

used 27 cycles (Opa-2), 33 cycles (Opa-3) or 31 cycles (Pilin-2). Each cycle consists of a<br />

denaturation step 85°C for 30 s and a hybridization, gap filling, ligation step of 60°C for<br />

Opa-2 and Pilin-2 and 53°C for Opa-3) for 1 min. When developing a new LCR procedure<br />

it is necessary to empirically test different numbers of cycles to determine the optimum<br />

for detection and specificity for each primer pair.<br />

6. Detection is performed on 40 µl of each LCR reaction in the automated Abbott Imx,<br />

which uses anti-fluorescein-coated microparticles to capture the products of the gap LCR.


Ligase Chain Reaction 143<br />

If ligation has taken place, the biotin on the other end of the product is detected by antibiotin<br />

conjugated to alkaline phosphatase. Alkaline phosphatase is detected by adding<br />

methylumbelliferone phosphate that, when the phosphatase acts on the phosphate, yields a<br />

fluorescent product that is detected by the IMx. (36). The amount of fluorescence produced<br />

is proportional to the ligated oligonucleotides.<br />

7. Of the many methods used to detect ligation products, automated methods such as the<br />

IMx are the most labor efficient. Details and requirements regarding the actual use of<br />

automated systems for LCR detection are instrument-specific and must be performed with<br />

the manufacturers in order to ensure the validity and accuracy of the results.<br />

3.3. Abbott LCx Commercial Automated Neisseria gonorrhoeae<br />

Detection System<br />

The materials listed here and the procedures described in the subsequent section are<br />

based on what is needed for performance of the commercially sold Abbott LCx assay<br />

for detection of N. gonorrhoeae from urogenital swabs or voided urine from men or<br />

women. Due to the proprietary nature of the LCx technology, use of these specific<br />

materials, including specimen transport systems and oligonucleotide primers that must<br />

be purchased from the manufacturer in kit form is necessary for the successful performance<br />

of this type of LCR assay. Strict adherence to the manufacturer’s instructions for<br />

sample collection, storage, and processing is necessary for satisfactory results. The LCx<br />

kits for C. trachomatis and M. tuberculosis follow the same general principles. Some<br />

modifications are necessary for the M. tuberculosis assay due to the nature of the different<br />

specimen types (respiratory secretions vs urogenital swabs or urine). It is possible to<br />

perform the LCR assay using other commercially available ligases and methods for<br />

detection of amplicons as described earlier, but the principles of the Abbott LCx and the<br />

notes regarding its performance are generally relevant for any type of LCR assay.<br />

3.3.1. Specimen Collection<br />

1. 15–20 mL of first-void urine should be collected into a sterile plastic, preservative-free<br />

container. It is desirable to obtain specimens from patients who have not urinated within<br />

1 h prior to collection.<br />

2. Swab specimens should be collected using the LCx swab collection and transport kit. For<br />

endocervical specimens in females, excess cervical mucus should be removed prior to<br />

sampling using the large-tipped cleaning swab provided in the collection system. When<br />

sampling the cervix, the small-tipped swab should be inserted into the endocervix and<br />

rotated for 15–30 s to ensure adequate sampling. In males, the small-tipped swab should<br />

be inserted 2–4 mm into the urethral meatus and rotated for 3 to 5 s. Swabs are then<br />

inoculated into the transport tube, broken off at the score line and then the cap is screwed<br />

securely onto the tube.<br />

3.3.2. Specimen Storage<br />

1. Time and temperature conditions must be adhered to for storage and transport of specimens.<br />

Swabs in the transport system can be stored at 2–30°C. If more than 24 h will elapse<br />

before processing, the swabs should ideally be refrigerated at 2–8°C or frozen at –20°C<br />

or below.<br />

2. Urine specimens should be refrigerated immediately at 2–8°C and can be held at this<br />

temperature for up to 4 d before processing. If longer storage is necessary, swab or urine<br />

specimens can be frozen at –20°C or below for up to 60 d. Do not thaw urine until ready<br />

for testing.


144 Benjamin, Smith, and Waites<br />

3.3.3. Urine Specimen Preparation and Processing<br />

1. Allow urine specimen to completely thaw if frozen. Mix urine in the urine collection cup<br />

by swirling to resuspend any settled material.<br />

2. Using a pipet with aerosol barrier pipet tips, transfer 1 mL of mixed urine into the Urine<br />

Specimen Microfuge Tube from the Urine Specimen Preparation Kit.<br />

3. Centrifuge at >9000g for 15 min in a microcentrifuge.<br />

4. Using a fine-tipped, plastic disposable pipet, gently aspirate all of the urine supernatant. Be<br />

cautious not to contact or dislodge the pellet, which may be translucent. The time between<br />

centrifugation and removal of supernatant must not exceed 15 min.<br />

5. Using a pipettor with aerosol barrier pipet tips, add 1 mL of LCx Urine Specimen<br />

Resuspension Buffer. Close lid of microfuge tube and resuspend the pellet by vortexing<br />

until the pellet is resuspended.<br />

6. Secure tube closure with a cap lock until it clicks into place.<br />

7. Insert specimen tubes in preheated dry bath wells. After the temperature of the heat block<br />

is stabilized at 97°C, heat specimens for 15 min.<br />

8. Remove the specimen from the dry bath and allow to cool at room temperature for 15 min.<br />

Remove cap lock and discard.<br />

9. Pulse-centrifuge the processed urine specimen in a microcentrifuge for a minimum of<br />

10–15 s immediately before inoculating the LCx amplification vials.<br />

10. The amplification reagent level in the LCx amplification vial should measure approx<br />

two-thirds of the conical part of the vial. If necessary, the vial may be pulse centrifuged<br />

in a microcentrifuge for 10–15 s.<br />

11. Using a pipet with aerosol barrier pipet tips, add 100 µL of each processed urine specimen<br />

to the appropriately labeled LCx Amplification Vial, making sure each vial is securely<br />

closed and that only one at a time is open. The vial can now be transferred to Area 2 and<br />

placed immediately in the Thermal Cycler for amplification.<br />

3.3.4. Swab Specimen Preparation<br />

1. Allow specimen to completely thaw, if frozen.<br />

2. Insert specimen tubes in preheated dry bath wells. After the temperature of the heat block<br />

is stabilized at 97°C, heat specimens for 15 min. Failure to reach 97 + 2°C could limit<br />

release of the DNA in the specimen causing false negative results.<br />

3. Remove the specimen from the dry bath and allow to cool at room temperature for<br />

15 min.<br />

4. Unscrew cap and express swab along the side of the tube so that liquid drains back into the<br />

solution at the bottom of the tube. Discard swab and original closure, replacing with a new<br />

swab tube closure that is screwed on until it clicks into place.<br />

5. The amplification reagent level in the LCx amplification vial should measure approx<br />

two-thirds of the conical part of the vial.<br />

6. Using a pipet with aerosol barrier pipet tips, add 100 µL of each processed specimen to<br />

the appropriately labeled LCx Amplification Vial, making sure each vial is securely closed<br />

and that only one at a time is open. The vial can now be transferred to Area 2 and placed<br />

immediately in the Thermal Cycler for amplification.<br />

7. The LCx negative control and calibrator must be prepared in conjunction with specimens<br />

to be tested and run in duplicate with each carousel of clinical specimens.<br />

3.3.5. Quality Control Procedures<br />

1. Negative control and calibrator preparations must take place in the dedicated Specimen<br />

Preparation Area (Area 1).


Ligase Chain Reaction 145<br />

2. The LCx procedure requires that the negative control and the calibrator be run in duplicate<br />

with each carousel of specimen.<br />

3. The negative control and calibrator are activated by the addition of 100 µL of LCx<br />

activation reagent. It is important to make sure correct volume is added or the run may<br />

be invalid. After addition, the contents of the bottles are then recapped and vortexed for<br />

20 s. Each bottle of activated negative control or calibrator is designed to be used up to<br />

48 h if stored at 2–8°C.<br />

4. A positive control that monitors the entire assay procedure including the specimen<br />

processing step should be tested. A known positive urine specimen can be processed in<br />

parallel and tested with unknown specimens. The positive control should give a positive<br />

assay value (S/CO ratio >1.00). Each laboratory should establish a target value and limits<br />

from each control batch using statistical control rules. These target values and limits should<br />

be maintained throughout storage. Alternative choices for a positive control are type strains<br />

of the microorganism targeted in the assay, e.g., N. gonorrhoeae.<br />

3.3.6. LCx Amplification Procedure<br />

1. Turn the LCx Thermal Cycler on for at least 15 min prior to use.<br />

2. Collect all LCx amplification vials containing samples, negative control and calibrator<br />

from Area 1 and transfer to Area 2 for thermal cycling.<br />

3. LCx thermal cycling conditions should be edited to the following amplification parameters:<br />

93°C for 1 s, 59°C for 1 s, 62°C for 1 min, 10 s for 40 cycles.<br />

4. Place the amplification vials into the thermal cycler, and initiate run. After completion<br />

of the thermal cycler run, amplification product may remain at 15–30°C for up to 72 h<br />

prior to LCx detection.<br />

3.3.7. Detection of Amplification Product<br />

1. Refer to the LCx Analyzer Operations Manual for detailed instrument operation procedures.<br />

Before running the LCx Analyzer, check to see that LCx Inactivation Diluent (1) contains<br />

a minimum of 100 mL and the LCx System Diluent (2) contains a minimum of 250 mL.<br />

Remove the LCx Amplification Vials from the LCx Thermal Cycler.<br />

2. Place LCx Reaction Cells into a Carousel; lock the carousel.<br />

3. Pulse-centrifuge the LCx amplification vials in a microcentrifuge for 10–15 s before<br />

placing into the LCx reaction cells.<br />

4. Place the amplification vials into the LCx reaction cells in the following order: negative<br />

controls in positions 1 and 2, calibrators in positions 3 and 4, and specimens in the<br />

remaining positions.<br />

5. Place the carousel into the LCx Analyzer.<br />

6. Lock the amplification vial Retainer.<br />

7. Remove the LCx detection reagent Pack from 2–8°C storage, gently invert it five times,<br />

and open the reagent pack bottles in the numeric order: 1, 2, 3, 4.<br />

8. Look for any film that may have formed over the opening of the reagent bottles. If present<br />

burst the bubble.<br />

9. Place the LCx detection reagent pack into the LCx Analyzer.<br />

10. Press Assay, then sample management to log in samples for the run. Press RUN on the LCx<br />

Analyzer control panel. Final assay results will be printed in approx 60 min.<br />

11. Store the detection reagent pack at 2–8°C in original packaging, decontaminated with 1% v/v<br />

hypochlorite or separate from unopened LCx kits.<br />

12. Remove the Carousel, individually remove the LCx reaction cells, and dispose appropriately.<br />

Rinse the carousel with water after each use.


146 Benjamin, Smith, and Waites<br />

3.3.8. Calculation of Results<br />

1. N. gonorrhoeae. The LCx Assay uses MEIA detection on the LCx Analyzer to detect<br />

DNA. All calculations are performed automatically.<br />

2. The presence or absence is determined by relating the LCx Assay results for the specimen<br />

to the Cutoff value. The Cutoff value is the mean RATE (c/s/s) of the LCx calibrator<br />

duplicates multiplied by 0.25.<br />

3. Calculation of the Cutoff value:<br />

Cutoff value = 0.25 × (Mean of LCx Gonorrhea Calibrator RATES)<br />

The S/CO value is determined by calculating a ratio of the sample RATE to the Cutoff value<br />

S<br />

CO<br />

Sample RATE<br />

Cutoff Value<br />

3.3.9. Interpretation of Results<br />

N. gonorrhoeae<br />

S/CO Ratio Result Report<br />

>1.20 LCx positive N. gonorrhoeae DNA detected, and positive for<br />

N. gonorrhoeae by LCR amplification and MEIA<br />

detection.<br />


Ligase Chain Reaction 147<br />

of contamination.<br />

3. When using LCx tests for detection of microorganisms in urogenital specimens, the<br />

presence of excessive mucus or blood in the swab or urine specimen as well as use of<br />

feminine powder sprays can interfere with the assay and cause false negative results, thus<br />

proper specimen quality must be monitored.<br />

4. In addition to the usual quality controls, reproducibility of LCx assays can be monitored<br />

by repeating random samples on a frequent basis.<br />

5. Further quality control testing and recording is detailed for the LCx analyzer and<br />

thermocycler in the respective manuals for the equipment.<br />

6. Failure to reach 97°C before amplification could limit release of the DNA in the specimen<br />

causing false negative results.<br />

7. Test the processed urine specimen immediately, or store for up to 60 d at 2–8°C or –20°C<br />

prior to testing. If the processed urine specimen is stored frozen, it must be completely<br />

thawed prior to addition to the LCx Amplification Vial.<br />

8. The LCx Analyzer first verifies that the assay results of the negative controls and calibrator<br />

are within the specified ranges of the LCx assay parameters by comparing the assay results<br />

of the negative control and calibrator to the values listed in the assay parameters. A run is<br />

valid when the individual and average results are within the values listed for CAL HIGH,<br />

CAL LOW, CAL AVE HIGH, CAL AVE LOW, NEG LOW, NEG HIGH, NEG AVE HIGH,<br />

and NEG AVE LOW parameters in the LCx assay parameters.<br />

9. In the event of an invalid negative control or calibrator assay result, the assay results<br />

printout will identify the out-of-range result, the S/CO ratio of the specimen will not be<br />

calculated and a flag indicating an invalid result will occur in the NOTE column of the<br />

printout. Ensure the LCx Negative Controls and Calibrators are in the correct order on the<br />

MEIA carousel to avoid an invalid run.<br />

10. Environmental quality control screening is also recommended. The laboratory should be<br />

monitored for the presence of amplification product by saturating one of the small-tipped<br />

swabs in transport buffer, then using it to wipe the desired area, including equipment,<br />

and processing according to the LCx procedures. If a positive reaction is detected, the<br />

decontamination steps using 1% sodium hypochlorite followed by 70% ethanol should be<br />

performed. The operation manual should be consulted for decontaminating equipment.<br />

References<br />

1. Barany, F. (1991) Genetic disease detection and DNA amplification using cloned thermostable<br />

ligase. Proc. Natl. Acad. Sci. USA 88, 189–193.<br />

2. Jurinke, C., van den Boom, D., Jacob, A., Tang, K., Worl, R., and Koster, H. (1996)<br />

Analysis of ligase chain reaction products via matrix-assisted laser desorption/ionization<br />

time-of-flight-mass spectrometry. Anal. <strong>Bio</strong>chem. 237, 174–181.<br />

3. Luo, J., Bergstrom, D. E., and Barany, F. (1996) Improving the fidelity of Thermus<br />

thermophilus DNA ligase. Nucleic Acids Res. 24, 3071–3078.<br />

4. Khanna, M., Park, P., Zirvi, M., Cao, W., Picon, A., Day, J., et al. (1999) Multiplex<br />

PCR/LDR for detection of K-ras mutations in primary colon tumors. Oncogene 18, 27–38.<br />

5. Khanna, M., Cao, W., Zirvi, M., Paty, P., and Barany, F. (1999) Ligase detection reaction for<br />

identification of low abundance mutations. Clin. <strong>Bio</strong>chem. 32, 287–290.<br />

6. Zirvi, M., Bergstrom, D. E., Saurage, A. S., Hammer, R. P., and Barany, F. (1999) Improved<br />

fidelity of thermostable ligases for detection of microsatellite repeat sequences using<br />

nucleoside analogs. Nucleic Acids Res. 27, e41.<br />

7. Day, D. J., Speiser, P. W., White, P. C., and Barany, F. (1995) Detection of steroid<br />

21-hydroxylase alleles using gene-specific PCR and a multiplexed ligation detection<br />

reaction. Genomics 29, 152–162.


148 Benjamin, Smith, and Waites<br />

8. Reyes, A. A., Carrera, P., Cardillo, E., Ugozzoli, L., Lowery, J. D., Lin, C. I., et al. (1997)<br />

Ligase chain reaction assay for human mutations: the Sickle Cell by LCR assay. Clin.<br />

Chem. 43, 40– 44.<br />

9. Besmer, P., Miller, R. J., Caruthers, M. H., Kumar, A., Minamoto, K., Van de Sande, J. H.,<br />

et al. (1972) Studies on polynucleotides. CXVII. Hybridization of polydeoxynucleotides<br />

with tyrosine transfer RNA sequences to the r- strand of phi80psu + 3 DNA. J. Mol. <strong>Bio</strong>l.<br />

72, 503–522.<br />

10. Landegren, U., Kaiser, R., Sanders, J., and Hood, L. (1988) A ligase-mediated gene<br />

detection technique. Science 241, 1077–1080.<br />

11. Alves, A. M. and Carr, F. J. (1988) Dot blot detection of point mutations with adjacently<br />

hybridising synthetic oligonucleotide probes. Nucleic Acids Res. 16, 8723.<br />

12. Wu, D. Y. and Wallace, R. B. (1989) The ligation amplification reaction (LAR)—<br />

amplification of specific DNA sequences using sequential rounds of template-dependent<br />

ligation. Genomics 4, 560–569.<br />

13. Morre, S. A., van Valkengoed, I. G., de Jong, A., Boeke, A. J., van Eijk, J. T., Meijer, C.<br />

J., and van den Brule, A. J. (1999) Mailed, home-obtained urine specimens: a reliable<br />

screening approach for detecting asymptomatic Chlamydia trachomatis infections. J. Clin.<br />

Microbiol. 37, 976–980.<br />

14. Chernesky, M. A., Jang, D., Sellors, J., Luinstra, K., Chong, S., Castriciano, S., and Mahony,<br />

J. B. (1997) Urinary inhibitors of polymerase chain reaction and ligase chain reaction and<br />

testing of multiple specimens may contribute to lower assay sensitivities for diagnosing<br />

Chlamydia trachomatis infected women. Mol. Cell Probes 11, 243–249.<br />

15. Prchal, J. T. and Guan, Y. L. (1993) A novel clonality assay based on transcriptional analysis<br />

of the active X chromosome. Stem Cells 11 (Suppl. 1), 62–65.<br />

16. Marshall, R. L., Laffler, T. G., Cerney, M. B., Sustachek, J. C., Kratochvil, J. D., and<br />

Morgan, R. L. (1994) Detection of HCV RNA by the asymmetric gap ligase chain reaction.<br />

PCR Methods Appl. 4, 80–84.<br />

17. Lee, H. H., Chernesky, M. A., Schachter, J., Burczak, J. D., Andrews, W. W., Muldoon,<br />

S., et al. (1995) Diagnosis of Chlamydia trachomatis genitourinary infection in women by<br />

ligase chain reaction assay of urine. Lancet 345, 213–216.<br />

18. Dille, B. J., Butzen, C. C., and Birkenmeyer, L. G. (1993) Amplification of Chlamydia<br />

trachomatis DNA by ligase chain reaction. J. Clin. Microbiol. 31, 729–731.<br />

19. Winn-Deen, E. S., Batt, C. A., and Wiedmann, M. (1993) Non-radioactive detection of<br />

Mycobacterium tuberculosis LCR products in a microtitre plate format. Mol. Cell Probes<br />

7, 179–186.<br />

20. Gaydos, C. A., Howell, M. R., Quinn, T. C., Gaydos, J. C., and McKee, K. T., Jr. (1998)<br />

Use of ligase chain reaction with urine versus cervical culture for detection of Chlamydia<br />

trachomatis in an asymptomatic military population of pregnant and nonpregnant females<br />

attending Papanicolaou smear clinics. J. Clin. Microbiol. 36, 1300–1304.<br />

21. Berg, E. S., Anestad, G., Moi, H., Storvold, G., and Skaug, K. (1997) False-negative results<br />

of a ligase chain reaction assay to detect Chlamydia trachomatis due to inhibitors in urine.<br />

Eur. J. Clin. Microbiol. Infect. Dis. 16, 727–731.<br />

22. Leckie, G. W., Erickson, D. D., He, Q., Facey, I. E., Lin, B. C., Cao, J., and Halaka, F. G.<br />

(1998) Method for reduction of inhibition in a Mycobacterium tuberculosis-specific ligase<br />

chain reaction DNA amplification assay. J. Clin. Microbiol. 36, 764–767.<br />

23. Buimer, M., van Doornum, G. J., Ching, S., Peerbooms, P. G., Plier, P. K., Ram, D., and<br />

Lee, H. H. (1996) Detection of Chlamydia trachomatis and Neisseria gonorrhoeae by<br />

ligase chain reaction-based assays with clinical specimens from various sites: implications<br />

for diagnostic testing and screening. J. Clin. Microbiol. 34, 2395–2400.


Ligase Chain Reaction 149<br />

24. Stary, A. (1999) Correct samples for diagnostic tests in sexually transmitted diseases: which<br />

sample for which test? FEMS Immunol Med Microbiol 24, 455– 459.<br />

25. Stary, A., Najim, B., and Lee, H. H. (1997) Vulval swabs as alternative specimens for ligase<br />

chain reaction detection of genital chlamydial infection in women. J. Clin. Microbiol.<br />

35, 836–838.<br />

26. Schepetiuk, S., Kok, T., Martin, L., Waddell, R., and Higgins, G. (1997) Detection of<br />

Chlamydia trachomatis in urine samples by nucleic acid tests: comparison with culture and<br />

enzyme immunoassay of genital swab specimens. J. Clin. Microbiol. 35, 3355–3357.<br />

27. Dubuis, O., Gorgievski-Hrisoho, M., Germann, D., and Matter, L. (1997) Evaluation of<br />

2-SP transport medium for detection of Chlamydia trachomatis and Neisseria gonorrhoeae<br />

by two automated amplification systems and culture for chlamydia. J. Clin. Pathol. 50,<br />

947–950.<br />

28. Hook, E. W., III, Smith, K., Mullen, C., Stephens, J., Rinehardt, L., Pate, M. S., and Lee, H.<br />

H. (1997) Diagnosis of genitourinary Chlamydia trachomatis infections by using the ligase<br />

chain reaction on patient-obtained vaginal swabs. J. Clin. Microbiol. 35, 2133–2135.<br />

29. Stary, A., Schuh, E., Kerschbaumer, M., Gotz, B., and Lee, H. (1998) Performance of<br />

transcription-mediated amplification and ligase chain reaction assays for detection of<br />

chlamydial infection in urogenital samples obtained by invasive and noninvasive methods.<br />

J. Clin. Microbiol. 36, 2666–2670.<br />

30. Carroll, K. C., Aldeen, W. E., Morrison, M., Anderson, R., Lee, D., and Mottice, S. (1998)<br />

Evaluation of the Abbott LCx ligase chain reaction assay for detection of Chlamydia<br />

trachomatis and Neisseria gonorrhoeae in urine and genital swab specimens from a sexually<br />

transmitted disease clinic population. J. Clin. Microbiol. 36, 1630–1633.<br />

31. Garrino, M. G., Glupczynski, Y., Degraux, J., Nizet, H., and Delmee, M. (1999) Evaluation<br />

of the Abbott LCx Mycobacterium tuberculosis assay for direct detection of Mycobacterium<br />

tuberculosis complex in human samples. J. Clin. Microbiol. 37, 229–232.<br />

32. Moore, D. F. and Curry, J. I. (1998) Detection and identification of Mycobacterium<br />

tuberculosis directly from sputum sediments by ligase chain reaction. J. Clin. Microbiol.<br />

36, 1028–1031.<br />

33. Ausina, V., Gamboa, F., Gazapo, E., Manterola, J. M., Lonca, J., Matas, L., et al. (1997)<br />

Evaluation of the semiautomated Abbott LCx Mycobacterium tuberculosis assay for direct<br />

detection of Mycobacterium tuberculosis in respiratory specimens. J. Clin. Microbiol.<br />

35, 1996–2002.<br />

34. Lindbrathen, A., Gaustad, P., Hovig, B., and Tonjum, T. (1997) Direct detection of<br />

Mycobacterium tuberculosis complex in clinical samples from patients in Norway by ligase<br />

chain reaction. J. Clin. Microbiol. 35, 3248–353.<br />

35. Fadda, G., Ardito, F., Sanguinetti, M., Posteraro, B., Ortona, L., Chezzi, C., et al. (1998)<br />

Evaluation of the Abbott LCx Mycobacterium tuberculosis assay in comparison with culture<br />

methods in selected Italian patients. New Microbiol. 21, 97–103.<br />

36. Birkenmeyer, L. and Armstrong, A. S. (1992) Preliminary evaluation of the ligase chain reaction<br />

for specific detection of Neisseria gonorrhoeae. J. Clin. Microbiol. 30, 3089–3094.<br />

37. Wilson, V. L., Wei, Q., Wade, K. R., Chisa, M., Bailey, D., Kanstrup, C. M., et al. (1999)<br />

Needle-in-a-haystack detection and identification of base substitution mutations in human<br />

tissues. Mutat. Res. 406, 79–100.<br />

38. Abravaya, K., Carrino, J. J., Muldoon, S., and Lee, H. H. (1995) Detection of point<br />

mutations with a modified ligase chain reaction (Gap-LCR). Nucleic Acids Res. 23,<br />

675–682.


150 Benjamin, Smith, and Waites


Nested RT-PCR in a Single Closed Tube 151<br />

26<br />

Nested RT-PCR in a Single Closed Tube<br />

Antonio Olmos, Olga Esteban, Edson Bertolini, and Mariano Cambra<br />

1. Introduction<br />

There are several methods now in widespread use for detecting and characterizing<br />

specific RNA targets. These methods include in situ hybridization, Northern blotting,<br />

dot or slot blot, RNase protection assay, and reverse transcription coupled to polymerase<br />

chain reaction (RT-PCR). However, when the amount of RNA target is limited or<br />

restricted in its cellular or tissue distribution, the extreme sensitivity of the PCR<br />

allows the detection of minute quantities of RNA when coupled to an initial step that<br />

converts single-strand RNA to cDNA (1–4). Nevertheless, when RT-PCR is applied<br />

for diagnostic purposes, the sensitivity usually afforded by this technique in routine<br />

detection tests is similar, or only slightly higher, to conventional enzyme-linked<br />

immunosorbent assay or hybridization techniques. This is frequently observed when<br />

dealing with poor-quality samples containing inhibitors of RT-PCR. The presence of different<br />

components of plant or animal origin, as well as specific RT-PCR conditions, may<br />

inhibit the reverse transcription and amplification by a number of mechanisms (5).<br />

Approaches have been developed to overcome these problems, including a previous<br />

immunocapture phase (6–8), the immobilization of RNA targets on plastic surfaces (9) or<br />

on paper membranes by a direct printing or squashing of the sample (10–12), preparation<br />

of crude extracts in simple buffers and subsequent dilution in sterile water, and other<br />

different protocols to prepare RNA targets free from interfering substances. Moreover,<br />

there are interesting alternatives to crude extracts or total nucleic acid preparations based<br />

on the use of commercially available resin- (13) or silica- (14) based kits.<br />

Sensitivity and specificity problems associated with RT-PCR may be overcome by<br />

using nested RT-PCR. The process is based on two consecutive rounds of amplification<br />

(15,16), the first round being an RT-PCR and the second a conventional PCR. The<br />

RT-PCR is performed using a pair of external primers. The 3′ end external primer (Pe1<br />

in Fig. 1) is used for RT reaction and the 3′ and 5′ end (Pe2 in Fig. 1) primers are<br />

used to perform the first PCR. The resulting amplification product is transferred to<br />

another Eppendorf tube containing a second pair of nested primers (nested PCR)<br />

that are internal to the initial pair. Alternatively, one of the external primers and a<br />

single nested primer (heminested PCR) can also be used. The larger amplification<br />

fragment produced during the first reaction is used as a target for the second (nested or<br />

heminested) PCR. The concept is illustrated in Fig. 1.<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

151


152 Olmos et al.<br />

Fig. 1. RT-heminested/nested-PCR scheme. Phase I (reverse transcription and first PCR):<br />

RNA target is copied in a cDNA form by the reverse transcriptase that uses the 3′ end external<br />

oligo (Pe1) as the reaction primer. Subsequently, the first PCR is performed using Pe1 and the<br />

5′ end external primer (Pe2) to produce as intermediate product that will act as the target in the<br />

second PCR amplification. Phase II (heminested or nested-PCR): a second PCR using a 5′ end<br />

internal primer (in the example Pi3) combined with one of the external primers (in the example<br />

Pe1; heminested-PCR) or two internal primers, the 3′ end internal primer (Pi1), and the 5′<br />

end internal primer (Pi2; nested PCR). Note that the internal primers could be used for typing<br />

purposes of the larger fragment amplified in the first reaction.


Nested RT-PCR in a Single Closed Tube 153<br />

Fig. 2. Compartmentalized Eppendorf tube that allows capture (if necessary), reverse<br />

transcription, and nested PCR in a single closed tube. Cocktail A for the reverse transcription<br />

and first PCR mix is added to the bottom of the tube. Cocktail B for the second amplification<br />

mix is added into the pipet tip cone.<br />

Several reports illustrate the potential of heminested and nested PCR in different<br />

fields (11,12,17–19). However, the use of two rounds of amplification in different<br />

tubes enhances the risk of contamination, especially when the method is used on a<br />

large scale. To prevent this problem, some authors proposed single-tube nested PCR<br />

protocols (20–22). However, a limitation to these approaches of nested RT-PCR is the<br />

need to accurately establish the ratio between an external and internal pair of primers<br />

and the use of limiting amounts of external primers to avoid its interference during the<br />

second amplification. Moreover, the external primers must be designed to anneal at<br />

a higher temperature than the internal primers. Figure 2 illustrates a simple device<br />

based on the use of a compartmentalized Eppendorf tube (Spanish patent P9801642<br />

of July 31, 1998) that allows RT reaction and nested PCR in a single closed tube in<br />

one manipulation. A small cone (the end of a standard 200-µL plastic pipet tip) is<br />

introduced into a 0.5-mL PCR tube allowing for the physical separation of the two<br />

different PCR cocktails in the same Eppendorf tube. The RT-PCR mix containing the<br />

external primers is added to the bottom of the tube and the PCR mix for the second<br />

amplification (heminested or nested) is added into the cone, where it remains as the<br />

result of capillary action (see Note 1). After RT-PCR, the Eppendorf tube is centrifuged<br />

to mix the products of the first reaction with the cocktail containing the internal primers.<br />

After this second round of PCR, the tube is finally opened to analyze the amplicons<br />

produced. The final result of this nested or heminested RT-PCR is a yield at least 100<br />

times higher than a conventional RT-PCR.<br />

The main advantages of this new approach of nested RT-PCR are the high sensitivity<br />

afforded without risk of contamination and the possibility of using external primers<br />

with the lowest annealing temperature and internal primers with the highest, in contrast<br />

to previously described protocols.


154 Olmos et al.<br />

This nested RT-PCR protocol coupled with a preparation of squashed or printed<br />

samples on paper (10) would allow the detection of RNA targets from a number of<br />

viruses in individual insect vectors, as well as in plant materials, animal fluids, or<br />

tissues. The increased sensitivity provided with this method permits the amplification of<br />

RNA targets from individual viruliferous aphids carrying stylet-borne (nonpersistent)<br />

and semipersistent plant viruses (11,12) without the need of a preliminary purification<br />

of nucleic acids.<br />

Conditions for multiplex nested RT-PCR can be easily established using the device<br />

described in Fig. 2. The simultaneous amplification of four RNA viruses and a<br />

bacterium from olive trees by multiplex nested RT-PCR without interference among<br />

primers has been successfully achieved in our laboratory (23). This system can include<br />

the use, if necessary (see Note 2), of an immunocapture (IC) phase in the same tube<br />

(IC-nested RT-PCR) (12). Recently, a new technique (called co-operational PCR/<br />

Co-PCR) for amplification of nucleic acids targets, based on a simple tetraprimer<br />

reaction has been described, with a sensitivity similiar to nested RT-PCR (24).<br />

2. Materials<br />

1. Eppendorf tubes (see Notes 2 and 3) (Cultek, Thermowell tubes cat no: 6530).<br />

2. Sterile 200-µL pipet tip cones (see Fig. 2 and Notes 1 and 4)(Daslab, cat. no: 16-2001).<br />

3. 10× RT-PCR buffer: 500 mM KCl, 100 mM Tris-HCl (pH 9.0 at 25°C), 1% Triton X-100<br />

(supplied with Taq DNA polymerase).<br />

4. Triton X-100 (Merck, Art. 8603).<br />

5. DMSO (Sigma, cat. no. D8418).<br />

6. MgCl 2 (25 mM; supplied with Taq DNA polymerase).<br />

7. dNTPs (5 mM, Pharmacia, cat. no. 27-2035-02).<br />

8. AMV reverse transcriptase (Pharmacia, cat. no. M5108).<br />

9. Taq DNA polymerase (Promega, cat. no. M1865).<br />

10. Primers for citrus tristeza virus (CTV) detection (see Subheading 3.2., step 1a):<br />

External primers (100 µM)<br />

• PEX1 (5′ TAA ACA ACA CAC ACT CTA AGG 3′)<br />

• PEX2 (5′ CAT CTG ATT GAA GTG GAC 3′)<br />

Internal primers (100 µM)<br />

• PIN1 (5′ GGT TCA CGC ATA CGT TAA GCC TCA CTT 3′)<br />

Primers for plum pox virus (PPV) detection (see Subheading 3.2., step 1b):<br />

External primers (100 µM<br />

• P10 (5′ GAG AAA AGG ATG CTA ACA GGA 3′)<br />

• P20 (5′ AAA GCA TAC ATG CCA AGG TA 3′)<br />

Internal primers (100 µM)<br />

• P1 (5′ ACC GAG ACC ACT ACA CTC CC 3′)<br />

• P2 (5′ CAG ACT ACA GCC TCG CCA GA 3′)<br />

11. Micropipets (Gilson, pipetman P10, P20, P100, P200).<br />

12. Microfuge (Heraeus, <strong>Bio</strong>fuge PICO).<br />

13. Thermal cycling machine with a heated lid (Techne, PHC-3).<br />

14. Electrophoresis system (<strong>Bio</strong>–Rad, sub-cell gt/powerpac 300 power supply system Cat.<br />

No. 165-4350).<br />

15. Ethidium bromide (10 µg/mL in water; AppliChem, A1152,0025).<br />

16. Transilluminator (TDI, Sepctroline Model TC-312A 312 nm UV).


Nested RT-PCR in a Single Closed Tube 155<br />

Table 1<br />

Volume and Concentration of Reactives for CTV Detection by Nested RT-PCR<br />

Ingredients Cocktail A (RT-PCR) Cocktail B (nested-PCR)<br />

10X RT-PCR buffer to 13.00 µl to 11.00 µl<br />

MgCl 2 (25 mM) to 13.60 µl –<br />

dNTPs (5mM each) to 12.25 µl –<br />

Triton ® X-100 (4%) to 12.50 µl –<br />

3′ external primer (100 µM) to 10.15 µl 3′ internal primer (100 µM) to 10.80 µl<br />

(Pe1 in Fig. 1) (Pi1 in Fig. 1)<br />

5′ external primer (100 µM) to 10.15 µl 5′ internal primer (100 µM) to 10.80 µl<br />

(Pe2 in Fig. 1) (Pi2 in Fig. 1)<br />

DMSO (100%) to 11.50 µl –<br />

H 2 O to 30.00 µl to 10.00 µl<br />

AMV (10 U/µl) to 10.20 µl –<br />

Taq Pol (5 U/µl) to 10.20 µl –<br />

3. Methods<br />

3.1. Primer Design (see Note 5)<br />

Sequenced regions of each RNA target can be recovered using the Nucleotide<br />

Sequence Search program located in the Entrez Browser program provided by the<br />

National Center for <strong>Bio</strong>technology Information (NCBI) (http://www3.ncbi.nlm.nih.gov/<br />

Entrez; Bethesda, MD). Conserved regions for each target can be studied using the<br />

similarity search Advanced BLAST 2.0, with the blast program designed to support<br />

analysis of nucleotides (http://www3.ncbi.nlm.nih.gov/blast/blast.cgi?Jform=1) (25).<br />

The alignment view could be performed as master-slave with identities to analyze<br />

significant nucleotide homologies in the molecular data retrieved from NCBI’s<br />

integrated databases, GenBank, EMBL, and DDBJ. Specific nucleotide regions should<br />

be selected. Different specific primers with similar annealing temperature can be<br />

subsequently designed for the target of interest, based on Oligo software utility<br />

(http://www.lifescience-software.com/oligo.htm; LRS, Long Lake, MN). By using<br />

this methodology, appropriate PCR primers to different RNA targets can be easily<br />

and properly designed.<br />

3.2. Nested PCR Product Generation<br />

1. RT-PCR and nested PCR cocktail preparation (see Fig. 2).<br />

a. When the annealing temperatures between external and internal primers are very<br />

different (external 45°C and internal 60°C) and the external primers do not anneal at<br />

60°C (for example, CTV detection), see Table 1.<br />

b. When the annealing temperatures of the external and internal primers are similar<br />

(external 50°C and internal 60°C) and the external primers anneal at 60°C (for example,<br />

PPV detection), see Table 2.<br />

2. Add 30 µL of cocktail A for reverse transcription and external amplification to the bottom<br />

of an 0.5-mL Eppendorf tube.<br />

3. Introduce the small plastic cone (see Fig. 2 and Notes 1 and 4) and add 10 µL of cocktail<br />

B into the cone.


156 Olmos et al.<br />

Table 2<br />

Volume and Concentration of Reactives for PPV Detection by Nested RT-PCR<br />

Ingredients Cocktail A (RT-PCR) Cocktail B (nested-PCR)<br />

10X RT-PCR buffer to 13.00 µl to 11.00 µl<br />

MgCl 2 (25 mM) to 13.60 µl –<br />

dNTPs (5 mM each) to 12.25 µl –<br />

Triton ® X-100 (4%) to 12.50 µl –<br />

3′ external primer (1 µM) to 13.00 µl 3′ internal primer (100 µM) to 10.80 µl<br />

(Pe1 in Fig. 1) (Pi1 in Fig. 1)<br />

5′ external primer (1 µM) to 13.00 µl 5′ internal primer (100 µM) to 10.80 µl<br />

(Pe2 in Fig. 1) (Pi2 in Fig. 1)<br />

DMSO to 11.50 µl –<br />

H 2 O to 30.00 µl to 10.00 µl<br />

AMV (10 U/µl) to 10.20 µl –<br />

Taq Pol (5 U/µl) to 10.20 µl –<br />

4. RT-PCR (see Note 6): 42°C for 45 min (reverse transcription), 94°C for 2 min (denaturation<br />

and reverse transcriptase inactivation), 20 to 25 cycles (92°C for 30 s [denaturation], 45 or<br />

50°C (see Subheading 3.2., step 1) for 30 s [annealing], and 72°C for 1 min [extension],<br />

and final elongation at 72°C for 10 min.<br />

5. Vortex the Eppendorf tube and centrifuge (pulse at 6000g for 2 s) to mix the second<br />

cocktail with the RT-PCR products.<br />

6. Nested-PCR (see Notes 7–9): 35 to 40 cycles (92°C for 30 s [denaturation], 60°C [see<br />

Subheading 3.2., step 1] for 30 s [annealing], and 72°C for 1 min [extension] and final<br />

elongation at 72°C for 10 min.<br />

7. Electrophoresis (see Fig. 3): Load 10 µL of amplification products onto a 3% agarose<br />

gel in 0.5× TAE and perform electrophoresis at 100 V for 30 min. Stain the agarose<br />

gel with ethidium bromide (0.5 µg/mL) for 15 min and visualize the amplicons under<br />

ultraviolet light.<br />

4. Notes<br />

1. The nested and heminested RT-PCR in a single closed tube described in this protocol<br />

is based on the use of a simple device (see Fig. 2). Other patented compartmentalized<br />

Eppendorf tubes with pockets could be commercialized by the industry in the near future,<br />

consequently avoiding the need to prepare cones. Accidental flow of the second PCR mix<br />

from the tip device might be caused by an incorrect manipulation of the device or by the<br />

use of pipet tips wider than standard. This can be solved by closing (heating) the tip end.<br />

In this case, after the RT-PCR it will be necessary to invert the tube, vortex to mix the<br />

reagents, and centrifuge to collect all components in the bottom of the tube.<br />

2. A previous capture step (immunocapture or print capture) improves the detection of some<br />

RNA targets by RT-PCR. Conventional immunocapture is usually performed by pre-coating<br />

Eppendorf tubes with 100 µL of carbonate buffer (pH 9.6) containing immunoglobulins<br />

(2 µg/mL) of a known specificity (6–8). Normal immunoglobulins from nonimmunized<br />

rabbits, bovine serum albumin, or skimmed milk can be also successfully used in this capture<br />

phase. However, because of the high sensitivity of nested or heminested RT-PCR, this step<br />

can usually be omitted. In this case, the detection of RNA targets can be performed by direct<br />

incubation of 100 µL of the sample (crude extract or tissue print Triton X-100 extracts from


Nested RT-PCR in a Single Closed Tube 157<br />

Fig. 3. Analysis of RT-nested-PCR products by agarose gel electrophoresis. (A) Detection<br />

of PPV targets from infected plant material. Lane 1: 100 bp molecular weight markers (Gibco<br />

BRL). Lanes 2 through 5: amplification products (243 bp) obtained from PPV-infected GF305<br />

peach seedlings. Lanes 6 and 7: samples from healthy GF305 peach seedlings. (B) Detection of<br />

CTV targets from infected aphid vectors. Lane 1: pUC19DNA/MspI(HpaII) molecular weight<br />

markers 23 (MBI Fermentas). Lanes 2 through 6: amplification products (131 bp) obtained from<br />

individual Aphis gossypii fed on CTV-infected Washington Navel sweet orange. Lane 7: Sample<br />

from Aphis gossypii fed on healthy Washington Navel sweet orange.<br />

paper; ref. 10) for 3 h at 37°C or overnight at 4°C in an uncoated Eppendorf tube. Before<br />

RT-PCR, the tube is washed twice with 150 µL of phosphate-buffered saline–0.05% Tween<br />

20 (washing buffer). However, it may be useful to assay the ability of different plastic tubes<br />

to trap RNA targets in order to select the most appropriate for RT-PCR.<br />

3. There are several commercially available thin-walled Eppendorf tubes for PCR. The<br />

protocol describes the use of 0.5-mL tubes because they are more easily compartmentalized<br />

with the end of a pipet tip (see Fig. 2).<br />

4. The reaction must be performed in a thermal cycler with a heated lid to prevent the<br />

requirement for oil-overlay during the cycling events. The selected thermal cycler must<br />

allow the maintenance of cocktail B (see Fig. 2) in the cone during the cycling. If<br />

evaporation occurs close the tip device as in Note 3 and/or adjust the lid temperature to 70<br />

to 80°C during the first reaction (RT-PCR).<br />

5. Primer design is critical to the success in amplifying RNA targets by conventional<br />

nested RT-PCR but it is even more critical when a multiplex nested strategy is used. The<br />

recommended primer design (see Subheading 3.1.) allows an optimal design of primers.<br />

The size of the amplified product should be small to ensure a good efficiency of the<br />

reaction and high sensitivity (26,27). Primers with a broad range of specificity must be<br />

designed from highly conserved genome sequences. Degenerate primers must be used for a<br />

universal detection of RNA targets belonging to a group. For characterization or typing of<br />

RNA targets, primers must be designed from discriminating or specific regions.<br />

6. The goal of the first round of PCR after RT is to obtain a yield of the amplification product<br />

just enough to ensure adequate target for the second (nested or heminested) amplification.<br />

For this reason, the number of cycles can be as low as 20 to 25.<br />

7. The goal of the second round of PCR (nested or heminested) is to achieve a high sensitivity<br />

and simultaneous specificity (if the method has been designed for typing RNA targets).<br />

The optimum number of cycles ranges from 35 to 40.


158 Olmos et al.<br />

8. The sensitivity of nested RT-PCR in a single closed tube can be compared with the<br />

sensitivity obtained by conventional RT-PCR using internal or external primers.<br />

9. The background of the nested RT-PCR must be reduced. Excessive cycling often results<br />

in the generation of multimeric PCR product that usually appears as a smear in the lane<br />

from the slot to the expected size of the amplified product in the agarose gel. Background<br />

also appears if the annealing temperature is not sufficient and/or the primer concentration<br />

is too high.<br />

Acknowledgments<br />

The authors wish to thank Dr. Mats Ohlin, University of Lund, Lund, Sweden, and<br />

Dr. Nuria Duran-Vila from IVIA, Valencia, Spain, for critical reading of the manuscript.<br />

The development of nested RT-PCR in a single closed tube has been funded by IVIA,<br />

INIA (SC98-060), CICYT (OLI96-2179), and EU-BIO96-0773 grants. O. Esteban<br />

and E. Bertolini were the recipients of PhD fellowships from Instituto Valenciano<br />

de Investigaciones Agarias and Agencia Española de Cooperación Internacional,<br />

respectively.<br />

References<br />

1. Hadidi, A., Levy, L., and Podleckis, E. V. (1995) Polymerase chain reaction technology in<br />

plant pathology, in Molecular Methods in Plant Pathology (Singh, R. P. and Singh, U. S.,<br />

eds.), CRC/Lewis Press, Boca Raton, FL, pp. 167–187.<br />

2. Candresse, T., Hammond, R. W., and Hadidi, A. (1996) Detection and identification of<br />

plant viruses and viroids using polymerase chain reaction (PCR), in Control of Plant Virus<br />

Diseases, APS Press, St. Paul, MN, pp. 399– 416.<br />

3. Singh, R. P. (1998) Reverse-transcription polymerase chain reaction for the detection of<br />

viruses from plants and aphids. J. Virol. Methods 74, 125–138.<br />

4. Hamoui, S., Benedetto, J. P., Garret, M., and Bonnet, J. (1994) Quantitation of mRNA<br />

species by RT-PCR on total mRNA population. PCR Methods Appl. 4, 160–166.<br />

5. Wilson, I. G. (1997) Inhibition and facilitation of nucleic acid amplification. Appl. Environ.<br />

Microbiol. 63, 3741–3751.<br />

6. Wetzel, T., Candresse, T., Macquaire, G., Ravelonandro, M., and Dunez, J. (1992) A<br />

highly sensitive immunocapture polymerase chain reaction method for plum pox potyvirus<br />

detection. J. Virol. Methods 39, 27–37.<br />

7. López-Moya, J. J., Cubero, J., López-Abella, D., and Díaz-Ruíz, J. R. (1992) Detection of<br />

cauliflower mosaic virus (CaMV) in single aphids by de polymerase chain reaction (PCR).<br />

J. Virol. Methods 37, 129–138.<br />

8. Nolasco, G., de Blas, C., Torres, V., and Ponz, F. (1993) A method combining immunocapture<br />

and PCR amplification in a microtiter plate for the detection of plant viruses and<br />

subviral pathogens. J. Virol. Methods 45, 201–218.<br />

9. Rowhani, A., Maningas, M. A., Lile, L. S., Daubert, S. D., and Golino, D. A. (1995)<br />

Development of a detection system for viruses of woody plants based on PCR analysis of<br />

immobilized virions. Phytopathology 85, 347–352.<br />

10. Olmos, A., Dasi, M. A., Candresse, T., and Cambra, M. (1996) Print-capture PCR: A simple<br />

and highly sensitive method for the detection of plum pox virus (PPV) in plant tissues.<br />

Nucleic Acids Res. 24, 2192–2193.<br />

11. Olmos, A., Cambra, M., Dasi, M. A., Candresse, T., Esteban, O., Gorris, M. T., and Asensio,<br />

M. (1997) Simultaneous detection and typing of plum pox potyvirus (PPV) isolates by<br />

Heminested-PCR and PCR-ELISA. J. Virol. Methods 68, 127–137.


Nested RT-PCR in a Single Closed Tube 159<br />

12. Olmos, A., Cambra, M., Esteban, O., Gorris, M. T., and Terrada, E. (1999) New device and<br />

method for capture, reverse transcription and nested-PCR in a single closed-tube. Nucleic<br />

Acids Res. 27, 1564–1565.<br />

13. Levy, L., Lee, I. M., and Hadidi, A. (1994) Simple and rapid preparation of infected plant<br />

tissue extracts for PCR amplification of virus, viroid and MLO nucleic acids. J. Virol.<br />

Methods 49, 295–304.<br />

14. Mackenzie, D. J., McLean, M. A., Mukerji, S., and Green, M. (1997) Improved RNA<br />

extraction from woody plants for the detection of viral pathogens by reverse transcriptionpolymerase<br />

chain reaction. Plant Dis. 81, 222–226.<br />

15. Simmonds, P., Balfe, P., Peutherer, J. F., Ludlam, C. A., Bishop, J. O., and Brown, A.<br />

J. (1990) Related articles human immunodeficiency virus-infected individuals contain<br />

provirus in small numbers of peripheral mononuclear cells and at low copy numbers.<br />

J. Virol. 64, 864–872.<br />

16. Porter-Jordan, K., Rosenberg, E. I., Keiser, J.F., Gross, J.D., Ross, A. M., Nasim, S., et al.<br />

(1990) Related articles nested polymerase chain reaction assay for the detection of<br />

cytomegalovirus overcomes false positives caused by contamination with fragmented<br />

DNA. J. Med. Virol. 30, 85–91.<br />

17. Arias, C. R., Garay, E., and Aznar, R. (1995) Nested PCR method for rapid and sensitive<br />

detection of Vibrio vulnificus in fish, sediments, and water. Appl. Environ. Microbiol.<br />

61, 3476–3478.<br />

18. Green, J., Henshilwood, K., Gallimore, C. I., Brown, D. W. G., and Lees, D. N. (1998)<br />

A nested reverse transcriptase PCR assay for detection of small round-structured viruses<br />

in environmentally contaminated molluscan shellfish. Appl. Environ. Microbiol. 64,<br />

858–863.<br />

19. Noyes, H., Reyburn, A. H., Bailey, J. W., and Smith, D. (1998) A nested-PCR-based<br />

schizodeme method for identifying Leishmania kinetoplast minicircle classes directly from<br />

clinical samples and its application to the study of the epidemiology of Leishmania tropica<br />

in Pakistan. J. Clin. Microbiol. 36, 2877–2881.<br />

20. Yourno, J. (1992) A method for nested PCR with single closed reaction tubes. PCR Methods<br />

Appl. 2, 60–65.<br />

21. Wolff, C., Hornschemeyer, D., Wolf, D., and Kleesiek, K. (1995) Single-tube nested PCR<br />

with room-temperature-stable reagents. PCR Methods Appl. 4, 376–379.<br />

22. Mutasa, E. S., Chwarszczynska, D. M., and Asher, M. J. C. (1996) Single-tube, nested PCR<br />

for the diagnosis of Polymyxa betae infection in sugar beet roots and colorimetric analysis<br />

of amplified products. Phytopathology 86, 493– 497.<br />

23. Bertolini, E., Olmos, A., López, M. M., and Cambra, M. (2003) Multiplex nested RT-PCR in<br />

a single tube for sensitive and simultaneous detection of four RNA viruses and Pseudomonas<br />

savastonoi pv savastanoi in olive trees. Phytopathology (in press).<br />

24. Olmos, A., Bertolini, E., and Cambra, M. (2002) Simultaneous and co-operational<br />

amplification (Co-PCR): a new concept for detection of plant viruses. J. Virol. Methods<br />

106, 51–59.<br />

25. Altschul, S. F., Madden, T. L., Schäfer, A. A., Zhang, J., Zhang, Z., Miller, W., et al. (1997)<br />

Gapped BLAST and PSI-BLAST: a new generation of protein database search programs.<br />

Nucleic Acids Res. 25, 3389–3402.<br />

26. Rosner, A., Maslenin, L., and Spiegel, S. (1997) The use of short and long PCR products<br />

for improved detection of prunus necrotic ringspot virus in woody plants. J. Virol. Methods<br />

67, 135–141.<br />

27. Singh, M. and Singh, R. P. (1997) Potato virus Y detection: sensitivity of RT-PCR depends<br />

on the size of fragment amplified. Can. J. Plant Pathol. 19, 149–155.


160 Olmos et al.


Direct PCR from Serum 161<br />

27<br />

Direct PCR from Serum<br />

Application to Viral Genome Detection<br />

Kenji Abe<br />

1. Introduction<br />

Nucleic acids used for polymerase chain reaction (PCR) assays usually are extracted<br />

by the phenol-chloroform method or an alternative rapid purification. The acidguanidinium<br />

thiocyanate-phenol-chloroform method for RNA extraction and proteinase<br />

K digestion-phenol-chloroform method for DNA extraction from serum samples are<br />

used widely for PCR assays (1,2), but usually at least 3 h are needed for this step.<br />

Detection of RNA is more difficult and complex than that of DNA, mainly because of<br />

the contaminating RNase and the need to conduct an additional reverse transcription<br />

(RT) step. Additionally, another problem is the high cost of the reagents for RNA<br />

extraction. Recently, we reported that viral RNA and DNA are readily amplified<br />

directly from serum without purification of nucleic acids by direct (RT) PCR (3). The<br />

method is sensitive because as little as 1 µL of the initial serum produces a clearly<br />

visible amplified fragment, and there is no difference in the sensitivity and stability<br />

between the direct PCR and conventional PCR assay. Interestingly, the results of the<br />

direct PCR are much better when 1 to 2 µL is used rather than 3 to 5 µL of serum.<br />

This inability to amplify RNA/DNA from serum may be to the result of the masking<br />

of the target RNA/DNA by coagulated proteins during the initial heat denaturation or<br />

the presence in the serum of inhibitors (probably such as lipoprotein) of the enzyme<br />

reaction. The sensitivity of the direct PCR assay allows detection of as few as one copy<br />

of viral genome. This technique not only eliminates the risk of cross contamination<br />

during nucleic acid extraction but also is cost and time saving. In this chapter, we<br />

describe the method of the direct (RT) PCR assay and apply it for detection of hepatitis<br />

B virus (HBV) DNA and hepatitis C virus (HCV) RNA in serum specimens.<br />

2. Materials<br />

1. Thermal cycler.<br />

2. AmpliTaq Gold DNA polymerase 5 U/µL (Perkin–Elmer).<br />

3. 10× reaction buffer containing 15 mM MgCl 2 (supplied with AmpliTaq Gold DNA<br />

polymerase kit).<br />

4. RNase inhibitor 40 U/µL (Promega).<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

161


162 Abe<br />

5. Moloney murine leukemia virus: 200 U(µL (M-MLV) reverse transcriptase (Gibco BRL).<br />

6. Primer sequences used for detection of HBV DNA and HCV RNA by direct (RT) PCR<br />

are listed in Table 1 (100 ng/µL).<br />

7. Deoxy nucleotide triphosphate (dNTP) mixture: 10 mM (Pharmacia <strong>Bio</strong>tech).<br />

8. Phosphate-buffered saline (PBS): Dissolve 8 g of NaCl, 0.2 g of KCl, 1.44 g of Na 2 HPO 4 ,<br />

and 0.24 g of KH 2 PO 4 in 800 mL of distilled H 2 O. Adjust the pH to 7.4 with HCl. Add<br />

H 2 O to 1 L. Dispense the solution into aliquots and sterilize them by autoclaving for<br />

20 min at 15 psi in. on liquid cycle. Store at room temperature.<br />

9. RNase-free H 2 O.<br />

10. 50-bp DNA ladder (Pharmacia <strong>Bio</strong>tech).<br />

3. Methods<br />

3.1 Pretreatment of Serum Samples (see Note 1)<br />

1. Serum samples (4 µL) are diluted with PBS to the final volume of 20 µL (15 dilution,<br />

see Note 1).<br />

2. The diluted serum samples are heated for 3 min at 95°C then cooled rapidly on ice for<br />

3 to 5 min.<br />

3. Five microliters of the heat-denatured diluted serum samples (= 1 µL of serum) are used<br />

as templates for the nested (RT) PCR.<br />

3.2. PCR Reaction<br />

1. Set up PCR mastermixes as follows.<br />

a. Reverse transcription and first PCR buffer (when performing DNA PCR omit RNase<br />

Inhibitor and M-MLV Reverse Transcriptase, see step 4).<br />

for RNA for DNA<br />

Ingredients µL µL Final concentration<br />

10× AmpliTaq Gold buffer 5 5 1× (1.5 mM MgCl 2 )<br />

(containing 15 mM MgCl 2 )<br />

10 mM dNTP mix 1 1 200 µM<br />

Sense primer (100 ng/µL) 1 1 100 ng<br />

Antisense primer (100 ng/µL) 1 1 100 ng<br />

AmpliTaq Gold DNA polymerase (5 U/µL) 0.4 0.4 2 U<br />

RNase inhibitor (40 U/µL) 0.25 None 10 U<br />

M-MLV reverse transcriptase (200 U/µL) 0.5 None 100 U<br />

RNase-free H 2 O 35.85 36.6<br />

Total 45 µL 45 µL<br />

b. Second PCR buffer<br />

Ingredients µL Final concentration<br />

10× AmpliTaq Gold buffer 5 1× (1.5 mM MgCl 2 )<br />

(containing 15 mM MgCl 2 )<br />

10 mM dNTP mix 1 200 µM<br />

Sense primer (100 ng/µL) 1 100 ng<br />

Antisense primer (100 ng/µL) 1 100 ng<br />

AmpliTaq Gold DNA polymerase (5 U/µL) 0.4 2 U<br />

RNase-free H 2 O 39.1<br />

Total 47.5 µL


163<br />

Table 1<br />

Primer Sequences Used for Nested PCR<br />

Primer Sequence Primer pair Product size<br />

For HBV DNA (X region)<br />

MD24 5′-TGCCAACTGGATCCTTCGCGGGACGTCCTT-3′ (nt 1392-1421) MD24/MD26 (1st) 233bp<br />

MD26 5′-GTTCACGGTGGTATAAATG-3′ (nt 1625-1607)<br />

HBx1 5′-GTCCCCTTCTTCATCTGCCGT-3′ (nt 1487-1507) HBx1/HBx2 (2nd) 117 bp<br />

HBx2 5′-ACGTGCAGAGGTGAAGCGAAG-3′ (nt 1604-1584)<br />

For HCV RNA (5′-untranslated region)<br />

19 5′-GCGACACTCCACCATAGAT-3′ (nt 2-20) 19/20 (1st) 329 bp<br />

20 5′-GCTCATGGTGCACGGTCTA-3′ (nt 330-312)<br />

21 5′-CTGTGAGGAACTACTGTCT-3′ (nt 28-46) 21/22 (2nd) 268 bp<br />

22 5′-ACTCGCAAGCACCCTATCA-3′ (nt 295-277)<br />

Nucleotide positions deduced from HBVadr4 (see ref. 4) for HBV and HC-J1 for HCV (see ref. 5)<br />

Direct PCR from Serum 163


164 Abe<br />

Table 2<br />

Detection Rate of Hepatitis Virus Nucleic Acids in Serum.<br />

Comparison of Conventional PCR and Direct PCR<br />

HBV DNA HCV RNA HGV RNA<br />

Conventional PCR Positive 121 56 32<br />

Direct PCR-Positive 121 51 30<br />

Agreement Rate 100% 91% 94%<br />

Fig. 1. Efficiency of serum volume and diluent as a factor in HCV-PCR reaction. A different<br />

amount of serum sample ranging in volume from 1 to 5 µL of serum was used. Each aliquot<br />

of serum, except for the 5 µL of volume, was diluted with PBS or H 2 O to the final volume of<br />

5 µL, respectively. M = 50-bp DNA ladder.<br />

2. Perform the first PCR reaction by adding 45 µL of mastermix 1 to each 5-µL sample of<br />

heat-inactivated serum from Subheading 3.1. Then, program the thermocycler to incubate<br />

the samples for 50 min at 37°C for the initial RT step and then preheat at 95°C for 10 min<br />

to activate AmpliTaq Gold followed by 50 cycles consisting of 94°C for 20 s, 55°C for 20 s,<br />

and 72°C for 30 s using a Thermal Cycler. RT-PCR is performed with a one-step method<br />

combined with cDNA synthesis reaction, followed by the PCR reaction in a single tube<br />

(see Note 2). That is, for RNA of HCV, the first PCR is combined with the RT step<br />

in the same tube containing 50 µL of a reaction buffer as shown above. To obtain an<br />

automatic hot-start reaction, we use the AmpliTaq Gold DNA polymerase instead of<br />

regular thermostable DNA polymerase.<br />

3. For the second reaction, 2.5 µL (1/20 volume) of the first PCR product is added to a tube<br />

containing second PCR buffer. The thermocycling for 50 cycles is performed as above, but<br />

omitting the initial 50 minute incubation at 37°C and the annealing temperature is raised<br />

to 60°C instead of 55°C for the second round of PCR.<br />

4. The PCR products are electrophoresed on a 2% agarose gels staining with ethidium<br />

bromide and evaluated under ultraviolet light. The sizes of PCR products are estimated<br />

according to the migration pattern of 50-bp DNA ladder.<br />

4. Notes<br />

1. Examination of the optimal serum quantity reveals that 1 to 2 µL of serum give a readily<br />

amplifiable band of HCV RNA but 3 µL of serum give a faint band and 4 to 5 µL of serum<br />

do not yield any band (Fig. 1). This result is much better when PBS is used to dilute serum<br />

than that obtained using H 2 O. It is similar even if this result is in the case of HBV.<br />

2. The whole process of the direct RT-PCR can be completed within 6 h by combination with<br />

the one-step amplification method and the second round PCR.<br />

3. Detection rate of HCV RNA and HBV DNA by direct PCR are consistent with the results<br />

obtained by conventional PCR (Fig. 2 and Table 2).


Direct PCR from Serum 165<br />

Fig. 2. Result of HCV-RNA detection in clinical serum samples. Comparison of conventional<br />

RT-PCR and direct RT-PCR.<br />

Fig. 3. An outline of direct (RT)-PCR.


166 Abe<br />

4. For HBV, nested PCR using the same PCR buffer for RT-PCR, but without reverse<br />

transcriptase and omitting cDNA synthesis step is performed.<br />

5. An outline of direct (RT)-PCR is shown in Fig. 3.<br />

References<br />

1. Chomczynski, P. and Sacchi, N. (1987) Single-step method of RNA isolation by acid<br />

guanidinium thiocyanate-phenol-chloroform extraction. Anal. <strong>Bio</strong>chem. 162, 156–159.<br />

2. Kunkel, L. M., Smith, K. D., Boyer, S. H., Borgaonkar, D. S., Wachtel, S. S., Miller, O. J.,<br />

et al. (1977) Analysis of human Y-chromosome-specific reiterated DNA in chromosome<br />

variants. Proc. Natl. Acad. Sci. USA 74, 1245–1249.<br />

3. Abe, K., Konomi, N., and Abdel-Hamid, M. (1998) Direct polymerase chain reaction<br />

for detection of hepatitis B, C and G virus genomes from serum without nucleic acid<br />

extraction—simple, rapid and highly sensitive method. Hepatol. Res. 13, 62–70.<br />

4. Fujiyama, A., Miyanchara, A., Nozaki, C., Yoneyama, T., Ohtoma, N., and Matsubara, K.<br />

(1983) Cloning and structural analyses of hepatitis B virus DNAs, subtype adr. Nucleic<br />

Acids Res. 11, 4601–4610.<br />

5. Okamoto, H., Okada, S., Sugiyama, Y., Yotsumoto, S., Tanaka, T., Yoshizawa, H., et al.<br />

(1990) the 5′-terminal sequence of the hepatitis C virus genome. Jpn J. Exp. Med. 60,<br />

167–177.


Long PCR 167<br />

28<br />

Long PCR Amplification of Large Fragments<br />

of Viral Genomes<br />

A Technical Overview<br />

Raymond Tellier, Jens Bukh, Suzanne U. Emerson, and Robert H. Purcell<br />

1. Introduction<br />

1.1. Long PCR<br />

The polymerase chain reaction (PCR) has become an essential and ubiquitous<br />

tool for biological research and laboratory diagnostic applications. Until recently,<br />

reliable and sensitive amplification of large templates (several kb) was difficult to<br />

achieve. However, in 1994, an important breakthrough was reported by Barnes (1).<br />

He hypothesized that a major obstacle to long PCR was the Taq DNA polymerase<br />

error rate, which causes mismatches that make elongation very inefficient. Many other<br />

thermostable DNA polymerases have a 3′ to 5′ exonuclease “proofreading” activity and<br />

a higher fidelity. However, the use of these polymerases alone does not reliably achieve<br />

long PCR, presumably because of excessive degradation of primers by the exonuclease<br />

activity (1). The processivity of the enzyme may also be a factor. Of note, the 3′ to 5′<br />

exonuclease activity alone is not a guarantee of high fidelity: Fidelity also depends<br />

on the degree of discrimination against misinsertion, the mismatch extension rate,<br />

and the rate of shuttling between polymerizing and proofreading modes (2). The<br />

breakthrough reported by Barnes consisted in performing PCR with a mixture of<br />

two DNA polymerases: a major component consisting of a highly processive DNA<br />

polymerase and a minor component consisting of a DNA polymerase with a 3′ to 5′<br />

exonuclease “proofreading” activity. With such enzyme mixes, reliable amplification<br />

of templates up to 35 kb in length was achieved (1). The greater fidelity of long PCR<br />

enzyme mixes, relative to Taq, has been demonstrated (1,2). Other modifications<br />

contribute to making long PCR possible, including optimization of the buffer and the<br />

thermal cycling conditions. It is especially important to address the drop in pH at high<br />

temperature observed with Tris-based buffers (1,3). This is significant because DNA<br />

depurination is enhanced both by high temperature and by low pH. Because a larger<br />

DNA template is more likely than a small template to sustain depurination at sites<br />

within its boundaries, long PCR is inherently more vulnerable to depurination (1,3).<br />

Interestingly, in contrast to 3′ to 5′ exonuclease-deficient DNA polymerases, the fidelity<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

167


168 Tellier et al.<br />

of Pfu (and therefore possibly of other proofreading DNA polymerases) actually<br />

increases with the pH (2). In general, long PCR works better with rapid temperature<br />

transition, which is addressed by using smaller reaction and mineral oil volumes,<br />

thin-wall PCR tubes, and rapid thermocyclers (1,3). Synthesis of amplicons of up to<br />

42 kb has now been reported (3).<br />

There are currently several commercially available long PCR kits, made from<br />

different enzymes and requiring somewhat different protocols. It is expected that new<br />

thermostable enzymes will be introduced. More research will be needed to determine<br />

which enzyme mix is superior for which application.<br />

1.2. Long RT-PCR<br />

Long PCR from mRNAs or from the genomes of RNA viruses obviously necessitates<br />

the synthesis of a sufficient amount of long cDNA. In some systems, such as plant<br />

viruses, it is possible to start with a large amount of RNA (4). For the more common<br />

scenario where the amount of RNA is limited, many have reported good results with the<br />

use of reverse-transcriptase enzymes that have been engineered to destroy the RNase H<br />

activity of the enzyme (5–15). In principle, this would prevent the premature destruction<br />

of the RNA template before the synthesis of a large cDNA copy. A potential problem<br />

with the synthesis of long cDNA is the presence of strong secondary structures in some<br />

template RNAs. To some extent, this problem can be addressed by raising the temperature<br />

as high as enzyme stability permits during reverse transcription. For example,<br />

the RNase H-deficient reverse transcriptase Superscript II is active at temperatures up<br />

to 50°C (although its optimal temperature is between 42 and 45°C; ref. 16). Another<br />

approach is to use the thermostable DNA polymerase rTth which in the presence of<br />

Mn 2+ has a reverse transcriptase activity at temperatures up to 70°C (17,18). A<br />

fundamental limitation with this approach is that the RNA hydrolysis catalyzed by<br />

divalent cations is accelerated by high temperature (17,19). Recent reports on the use of<br />

nucleic acid isostabilizers, such as betaine (20), and the thermal stabilization of enzymes<br />

by trehalose (18) may lead to further refinements in this area.<br />

2. Applications of Long PCR<br />

2.1. General Applications<br />

The availability of long PCR obviates some shortcomings of standard PCR: Not only<br />

does it speed up cloning or sequencing of genetic sequences of interest, but it extends<br />

the power of the PCR by enabling it to bridge large gaps of unknown or variable<br />

sequences between primers. As examples of the latter, there have been many reports of<br />

the use of long PCR to amplify introns of unknown sequence using primers targeting<br />

the adjacent exons (e.g., refs. 21–24), or to characterize transposons in Drosophila (25).<br />

Long PCR has also been used to amplify the complete mitochondrial genome (26) and<br />

large genomic fragments for the determination of deletions in genetic diseases (27,28)<br />

or gene fusions in cancer (29). Other applications in microbiology include bacterial<br />

typing (30,31) and amplification of Plasmodium falciparum DNA (32).<br />

2.2. Long PCR and Viruses<br />

Long PCR has been applied with great success to viral genomes. For example,<br />

amplification of full-length, near full-length, or large fragments of the genome of many<br />

DNA viruses has been achieved, such as for the hepatitis B virus (HBV) (8,33), the


Long PCR 169<br />

human papilloma virus (34), SV-40 (35), varicella-zoster virus (36), the proviruses<br />

of the simian foamy virus of chimpanzees (SFVcpz) (37), human T-cell leukemia<br />

virus 1 (HTLV-1) (38), and human immunodeficiency virus type 1 (HIV-1) (39–41)<br />

and type 2 (HIV-2) (42).<br />

Long RT-PCR for RNA viruses has also been successful. Large fragments of the viral<br />

genome have been amplified for the tick-borne encephalitis virus (6), the Norwalk-like<br />

viruses (12), the hepatitis C virus (HCV) (13), and the bovine torovirus (43). The near<br />

full-length genome was amplified for HCV (8,14) and the full-length genome for the<br />

potato virus Y (4), hepatitis A virus (HAV) (7,9), hepatitis E virus (44), poliovirus (11),<br />

and coxsackie B2 virus (10,15).<br />

For HAV, we have shown that RNA transcribed directly from the full-length amplicon<br />

is infectious (7,9). Furthermore, long PCR has greatly facilitated the construction<br />

of infectious clones or infectious cDNAs for HBV (33), SV-40 (35), SFVcpz (37),<br />

HIV-1 (41), tick borne encephalitis virus (6), HAV (9), coxsackie B2 (15), coxsackie<br />

B6 (45), HCV (46,47), and potato virus Y (4). These results confirm the high fidelity<br />

of long PCR.<br />

In addition to streamlining the cloning and sequencing of viral genomes, long PCR<br />

presents other advantages for virology. For example, many viral sequences are toxic to<br />

Escherichia coli, leading to selection bias in cloning procedures (48). This problem<br />

can be circumvented to a great extent by long PCR because large amplicons can be<br />

directly sequenced (or transcribed). Furthermore, direct sequencing of amplicons<br />

provides a certain protection against DNA polymerase mistakes during PCR: Unless<br />

these mistakes occur in the early cycles, at each position the majority of amplicons will<br />

have the correct nucleotide and the sequencing will provide the correct result, whereas<br />

cloning might select an amplicon with mistakes. However, one must remember<br />

that there are circumstances, such as when encountering homopolymeric stretches,<br />

tandem repeats, extreme AT or GC content, or strong secondary structures, in which<br />

polymerases can produce systematic mistakes (49–52).<br />

Long PCR is also of great interest for RNA viruses, which typically exist as<br />

“quasispecies,” a mixed population containing several variants. Not only does direct<br />

sequencing of the amplicon yield immediately the consensus sequence, but it has been<br />

proposed that long PCR can be used to preserve and propagate a representative DNA<br />

version of such a “quasispecies” (9,11). In fact, it has been shown that long PCR<br />

preserves the distribution of variants of poliovirus (11) and HIV-1 (53). In the latter<br />

study, however, a relatively high rate of recombination between templates was observed<br />

after performing PCR (53). Recombination was measured by cloning and sequencing,<br />

and thus this measurement is possibly affected by a selection bias from the cloning<br />

process. Nonetheless, because of this observation, and pending further research on<br />

this topic, the possibility of artifactual recombination should be kept in mind. Because<br />

the recombination is thought to occur as the result of incomplete transcripts acting<br />

as “mega primers” in subsequent PCR cycles (53,54), careful optimization of the<br />

elongation time during PCR may minimize this problem. The use of polymerases with<br />

high processivity may also diminish the occurrence of recombination (54).<br />

References<br />

1. Barnes, W. M. (1994) PCR amplification of up to 35-kb DNA with high fidelity and high<br />

yield from λ bacteriophage templates. Proc. Natl. Acad. Sci. USA 91, 2216–2220.


170 Tellier et al.<br />

2. Cline, J., Braman, J. C., and Hogrefe, H. H. (1996) PCR fidelity of Pfu DNA polymerase<br />

and other thermostable DNA polymerases. Nucleic Acids Res. 24, 3546–3551.<br />

3. Cheng, S., Fockle, C., Barnes, W. M., and Higuchi, R. (1994) Effective amplification of<br />

long targets from cloned inserts and human genomic DNA. Proc. Natl. Acad Sci. USA<br />

91, 5695–5699.<br />

4. Fakhfakh, H., Vilaine, F., Makni, M., and Robaglia, C. (1996) Cell-free cloning and biolistic<br />

inoculation of an infectious cDNA of potato virus Y. J. Gen. Virol. 77, 519–523.<br />

5. Nathan, M., Mertz, L. M., and Fox, D. K. (1995) Optimizing long RT-PCR. Focus 17,<br />

78–80<br />

6. Gritsun, T. S. and Gould, E. A. (1995) Infectious transcripts of tick-borne encephalitis<br />

virus, generated in days by RT-PCR. Virology 214, 611–618.<br />

7. Tellier, R., Bukh, J., Emerson, S. U., and Purcell, R. H. (1996) Amplification of the fulllength<br />

hepatitis A virus genome by long reverse transcription-PCR and transcription of<br />

infectious RNA directly from the amplicon. Proc. Natl. Acad. Sci. USA 93, 4370– 4373.<br />

(Erratum published in Proc. Natl. Acad. Sci. USA 1997;94:10,485)<br />

8. Tellier, R., Bukh, J., Emerson, S. U., Miller, R. H., and Purcell, R. H. (1996) Long PCR and<br />

its application to hepatitis viruses: amplification of hepatitis A, hepatitis B and hepatitis C<br />

virus genomes. J. Clin. Microbiol. 34, 3085–3091. (Erratum published in J. Clin. Microbiol.<br />

1997;35:2713).<br />

9. Tellier, R., Bukh, J., Emerson, S. U., and Purcell, R. H. (1997) Amplification of the<br />

full length hepatitis A virus by long RT-PCR and generation of infectious RNA directly<br />

from the amplicon, in Viral Hepatitis and Liver Disease Proceedings of the IX Triennial<br />

International Symposium on Viral Hepatitis and Liver Disease (Rizzetto, M., Purcell, R. H.,<br />

Gerin, J. L., and Verme G., eds.), Minerva Medica, Turin, pp. 48–50.<br />

10. Gow, J. W., McGill, M. M., Behan, W. M. H., and Behan, P. O. (1996) Long RT-PCR<br />

amplification of full-length enterovirus genome. <strong>Bio</strong>Techniques 20, 582–584.<br />

11. Chumakov, K. M. (1996) PCR engineering of viral quasispecies: A new method to preserve<br />

and manipulate genetic diversity of RNA virus populations J. Virol. 70, 7331–7334.<br />

12. Ando, T., Monroe, S. S., Noel, J. S., and Glass, R. I. (1997) A one tube method of reverse<br />

transcription-PCR to efficiently amplify a 3-kilobase region from the RNA polymerase<br />

gene to the poly(A) tail of small round-structured viruses (Norwalk-like viruses). J. Clin.<br />

Microbiol. 35, 570–577.<br />

13. Rispeter, K., Lu, M., Lechner, S., Zibert, A., and Roggendorf, M. (1997) Cloning and<br />

characterization of a complete open reading frame of the hepatitis C virus genome in only<br />

two cDNA fragments. J. Gen. Virol. 78, 2751–2759.<br />

14. Wang, L-F., Radkowski, M., Vargas, H., Rakela, J., and Laskus, T. (1997) Amplification and<br />

fusion of long fragments of hepatitis C virus genome. J. Virol. Methods 68, 217–223.<br />

15. Lindberg, A. M., Polacek, C., and Johansson, S. (1997) Amplification and cloning of<br />

complete enterovirus genomes by long distance PCR. J. Virol. Methods 65, 191–199.<br />

16. Gibco BRL (1994) Technical bulletin 18064-2. Gibco BRL, Gaithersburg, Md.<br />

17. Myers, T. W. and Gelfand, D. H. (1991) Reverse transcription and DNA amplification by a<br />

Thermus thermophilus DNA polymerase. <strong>Bio</strong>chemistry 30, 7661–7666.<br />

18. Carninci, P., Nishiyama, Y., Westover, A., Itoh, M., Nagaoka, S., Sasaki, N., et al. (1998)<br />

Thermostabilization and thermoactivation of thermolabile enzymes by trehalose and its<br />

application for the synthesis of full length cDNA. Proc. Natl. Acad. Sci. USA 95, 520–524.<br />

19. Lindahl, T. (1967) Irreversible heat inactivation of transfer ribonucleic acids. J. <strong>Bio</strong>l.<br />

Chem. 242, 1970–1973.<br />

20. Henke, W., Herdel, K., Jung, K., Schnorr, D., and Loening, S. A. (1997) Betaine improves<br />

the PCR amplification of GC-rich DNA sequences. Nucleic Acids Res. 25, 3957–3958.


Long PCR 171<br />

21. Gaedigk, R., Karges, W., Hui, M. F., Scherer, S. W., and Dosch, H.-M. (1996) Genomic<br />

organization and transcript analysis of ICAp69, a target antigen in diabetic autoimmunity.<br />

Genomics 38, 382–391.<br />

22. Purroy, J., Bisceglia, L., Calonge, M. J., Zelante, L., Testar, X., Zorzano, A., et al. (1996)<br />

Genomic structure and organization of the human rBAT gene (SLC3A1). Genomics 37,<br />

249–252.<br />

23. Nobile, C., Marchi, J., Nigro, V., Roberts, R. G., and Danieli, G. A. (1997) Exon-intron<br />

organization of the human dystrophin gene. Genomics 45, 421– 424.<br />

24. Kelley, P. M., Weston, M. D., Chen, Z.-Y, Orten, D. J., Hasson, T., Overbeck, L. D., et al.<br />

(1997) The genomic structure of the gene defective in Usher syndrome type 1b (MYO7A).<br />

Genomics 40, 73–79.<br />

25. Gordadze, A. V. and Beněs, H. (1996) Long PCR-based technique for detection of transposon<br />

insertions in and around cloned genes of Drosophila melanogaster. <strong>Bio</strong>Techniques<br />

21, 1062–1066.<br />

26. Li, Y.-Y., Hengstenberg, C., and Maisch, B.(1995) Whole mitochondrial genome amplification<br />

reveals basal level multiple deletions in mtDNA of patients with dilated cardiomyopathy.<br />

<strong>Bio</strong>chem. <strong>Bio</strong>phys. Res. Commun. 10, 211–218.<br />

27. Bochmann, H., Gehrisch, S., and Jaross, W. (1996) Fast amplification of the low density<br />

lipoprotein receptor gene and detection of a large deletion by means of long polymerase<br />

chain reaction. Eur. J. Clin. Chem. Clin. <strong>Bio</strong>chem. 34, 955–959.<br />

28. Hećimović, S., Barišić, I., Müller, A., Pertković, I., Barić, I., Ligutić, et al. (1997) Expand<br />

long PCR for fragile X mutation detection. Clin. Genet. 52, 147–154.<br />

29. Waggott, W., Lo, Y.-M. D., Bastard, C., Gatter, K. C., Leroux, D., Mason, D. Y., et al.<br />

(1995) Detection of NPM-ALK DNA rearrangement in CD30 positive anaplastic large-cell<br />

lymphoma. Br. J. Haematol. 89, 905–907.<br />

30. Smith-Vaughan, H. C., Sriprakash, K. S., Mathews, J. D., and Kemp, D. J. (1995)<br />

Long PCR-ribotyping of nontypeable Haemophilus influenzae. J. Clin. Microbiol. 33,<br />

1192–1195.<br />

31. Hookey, J. V., Saunders, N. A., Clewley, J. P., Efstratiou, A., and George, R. C. (1996)<br />

Virulence regulon polymorphism in group A streptococci revealed by long PCR and implications<br />

for epidemiological and evolutionary studies. J. Med. Microbiol. 45, 285–293.<br />

32. Su, X.-Z., Wu, Y., Sifri, C. D., and Wellems, T. E. (1996) Reduced extension temperatures<br />

required for PCR amplification of extremely A+T– rich DNA. Nucleic Acids Res. 24,<br />

1574–1575.<br />

33. Günther, S., Li, B.-C., Miska S., Krüger, D. H. Meisel, H., and Will, H. (1995) A novel<br />

method for efficient amplification of whole hepatitis B virus genomes permits rapid<br />

functional analysis and reveals deletion mutants in immunosuppressed patients. J. Virol.<br />

69, 5437–5444.<br />

34. Stewart, A.-C. M., Gravitt, P. E., Cheng, S., and Wheeler, C. M. (1995) Generation of<br />

entire human papillomavirus genomes by long PCR: frequency of errors produced during<br />

amplification. Genome Res. 5, 79–88.<br />

35. Lednicky, J. A., Jafar, S., Wong, C., and Butel, J. S. (1997) High-fidelity PCR amplification<br />

of infectious copies of the complete simian virus 40 genome from plasmids and virusinfected<br />

cell lysates. Gene 184, 189–195.<br />

36. Takayama, M., Takayama, N., Inoue, N., and Kameoka, Y. (1996) Application of long PCR<br />

method to identification of variations in nucleotide sequences among Varicella-Zoster virus<br />

isolates. J. Clin. Microbiol. 34, 2869–2874.<br />

37. Herchenröder, O., Turek, R., Neumann-Haefelin, D., Rethwilm, A., and Schneider, J. (1995)<br />

Infectious proviral clones of chimpanzee foamy virus(SFVcpz) generated by long PCR<br />

reveal close functional relatedness to human foamy virus. Virology 214, 685–689.


172 Tellier et al.<br />

38. Tamiya, S., Matsuoka, M., Etoh, K.-I., Watanabe, T., Kamihira, S., Yamaguchi K., et al.<br />

(1996) Two types of defective human T-lymphotropic virus type 1 provirus in adult T-cell<br />

leukemia. Blood 88, 3065–3073.<br />

39. Salminen, M. O., Koch, C., Sanders-Buell, E., Ehrenberg, P. K., Michael, N. L., Carr, J. K.,<br />

et al. (1995) Recovery of virtually full-length HIV-1 provirus of diverse subtypes from<br />

primary virus cultures using the polymerase chain reaction. Virology 213, 80–86.<br />

40. Salminen, M. O., Johansson, B., Sönnerborg, A., Ayehunie, S., Gotte, D., Leinikki, P., et al.<br />

(1996) Full-length sequence of an Ethiopian human immunodeficiency virus type 1 (HIV-1)<br />

isolate of genetic subtype C. Aids Res. Hum. Retroviruses 12, 1329–1339.<br />

41. Gao, F., Robertson, D. L., Carruthers, C. D., Morrison, S. G., Jian, B., Chen, Y., et al. (1998)<br />

A comprehensive panel of near-full-length clones and reference sequences for non-subtype<br />

B isolates of human immunodeficiency virus type 1. J. Virol. 72, 5680–5698.<br />

42. Damond, F., Loussert-Ajaka, I., Apetrei, C., Descamps, D., Souquière, S., Leprêtre, A.,<br />

et al. (1998) Highly sensitive method for amplification of human immunodeficiency virus<br />

type 2 DNA. J. Clin. Microbiol. 36, 809–811.<br />

43. Duckmanton, L. M., Tellier, R., Liu, P., and Petric, M. (1998) Bovine torovirus: sequencing<br />

of the structural genes and expression of the nucleocapsid protein of Breda virus. Virus<br />

Res. 58, 83–96.<br />

44. Donati, M. C., Fagan, E. A., and Harrison, T. J. (1996) Sequence analysis of full length<br />

HEV clone derived directly from human liver in fulminant hepatitis E, in Viral Hepatitis<br />

and Liver Disease, Proceedings of the IX Triennial International Symposium on Viral<br />

Hepatitis and Liver Disease (Rizzetto, M., Purcell, R. H., Gerin, J., and Verme, G., eds.),<br />

Minerva Medica, Turin, pp. 313–316.<br />

45. Martino, T. A., Tellier, R., Petric, M., Irwin, D. M., Afshar A., and Liu P. (1999) The<br />

complete consensus sequence of coxsackievirus B6 and generation of infectious clones by<br />

long RT-PCR. Virus Res. 64, 77–86.<br />

46. Yanagi, M., Purcell, R. H., Emerson S. U., and Bukh, J. (1997) Transcripts from a single<br />

full-length cDNA clone of hepatitis C virus are infectious when directly transfected into the<br />

liver of a chimpanzee. Proc. Natl. Acad. Sci. USA 94, 8738–8743.<br />

47. Yanagi, M., St. Claire, M., Shapiro, M., Emerson, S. U., Purcell., R. H., and Bukh, J. (1998)<br />

Transcripts of a chimeric cDNA clone of hepatitis C virus genotype 1b are infectious in<br />

vivo. Virology 244, 161–172.<br />

48. Forns, X., Bukh, J., Purcell, R. H., and Emerson, S. U. (1997) How Escherichia coli can bias<br />

the results of molecular cloning: preferential selection of defective genomes of hepatitis C<br />

virus during the cloning procedure. Proc. Natl. Acad. Sci. USA 94, 13,909–13,914.<br />

49. Loewen, P. C. and Switala, J. (1995) Template secondary structure can increase the error<br />

frequency of the DNA polymerase from Thermus aquaticus. Gene 164, 59–63.<br />

50. Wu, M. J., Chow, L. W., and Hsieh, M. (1998) Amplification of GAA/TTC triplet repeat in<br />

vitro: Preferential expansion of (TTC)n strand. <strong>Bio</strong>chem. <strong>Bio</strong>phys. Acta 1407, 155–162.<br />

51. Ji, J., Clegg, N. J., Peterson, K. R., Jackson, A. L., Laird, C. D., and Loeb, L. A. (1996) In<br />

vitro expansion of GGC:GCC repeats: identification of the preferred strand of expansion.<br />

Nucleic Acids Res. 24, 2835–2840.<br />

52. Kroutil, L. C., Register, K., Bebenek, K., and Kunkel, T. A. (1996) Exonucleolytic<br />

proofreading during replication of repetitive DNA. <strong>Bio</strong>chemistry 35, 1046–1053.<br />

53. Yang, Y. L., Wang, G., Dorman, K., and Kaplan, A.H. (1996) Long polymerase chain<br />

reaction amplification of heterogeneous HIV type 1 templates produces recombination at a<br />

relatively high frequency. Aids Res. and Hum. Retroviruses 12, 303–306.<br />

54. Saiki, R. K., Gelfand, D. H., Stoffel, S., Scharf, S. J., Higuchi, R., Horn, G. T., et al. (1988)<br />

Primer directed enzymatic amplification of DNA with a thermostable DNA polymerase.<br />

Science 239, 487– 491.


Long PCR Methodology 173<br />

29<br />

Long PCR Methodology<br />

Raymond Tellier, Jens Bukh, Suzanne U. Emerson, and Robert H. Purcell<br />

1. Introduction<br />

In this chapter, we detail protocols of long polymerase chain reaction (PCR) and<br />

long RT-PCR, which we have found to be versatile, sensitive, and straightforward to<br />

optimize. We have used these protocols with success on several different templates,<br />

including lambda phage DNA, HAV, HBV, HCV (1), torovirus (2), coxsackie B6<br />

virus (3), and human beta galactosidase mRNA (R. Tellier, unpublished data). The<br />

guanine–cytosine (GC) content of these templates varied from 37.8 to 58.8%. These<br />

protocols have also been used on genomic human DNA with success (4).<br />

The long PCR protocol we use is derived from the method described by Barnes<br />

(5) for KLA-16, a mixture of KlenTaq 1 and Pfu. We have replaced this mix by a<br />

commercially available mix, Advantage KlenTaq Polymerase mix (Clontech), a mix<br />

of KlenTaq 1, and Deep Vent. In our hands, this mix was found to produce more<br />

consistent results. KlenTaq 1 is an N-terminal deletion mutant of Taq, analogous to<br />

the Klenow fragment enzyme, and devoid of the 5′- 3′ exonuclease activity of Taq (5).<br />

Unexpectedly, KlenTaq 1 was also found to be more thermostable and slightly more<br />

accurate than Taq (5,6), as well as less sensitive to variation in Mg 2+ concentration<br />

and more processive than Taq (7).<br />

2. Materials<br />

1. Water free of RNase, DNAse, and proteinase (ddH 2 O). Aliquot and store at –20°C.<br />

2. DTT (100 mM, Promega or Gibco BRL). Store at –20°C.<br />

3. RNasin (20–40 U/µL; Promega). Store at –20°C.<br />

4. Superscript II reverse transcriptase (Gibco BRL/Life Technologies). Store at –20°C.<br />

5. 5× RT buffer (Gibco BRL). Store at –20°C.<br />

6. RNase H (1–4 U/µL; Gibco BRL). Store at –20°C.<br />

7. RNase T1 (900–3000 U/µL; Gibco BRL). Store at –20°C.<br />

8. 50× Advantage KlenTaq Polymerase Mix (Clontech). Store at –20°C(see Note 1).<br />

9. 10× KlenTaq PCR buffer (Clontech). Store at –20°C (see Note 2).<br />

10. dNTP mix (10 mM each). Prepared in ddH 2 O from 100 mM dNTP set (Pharmacia). Aliquot<br />

and store at –80°C.<br />

11. Mineral oil (molecular biology grade, Sigma).<br />

12. Primers. Aliquot at a concentration of 10 µM in ddH 2 O, store at –20°C (see Note 3).<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

173


174 Tellier et al.<br />

13. Thin-wall PCR tubes (Stratagene; see Note 4).<br />

14. Robocycler 40 (Stratagene; see Note 5).<br />

3. Methods: Long PCR<br />

3.1. Long PCR of DNA<br />

1. Components of the PCR are prepared in a master mix. For each reaction use the following:<br />

10× KlenTaq PCR buffer (5 µL); 1.25 µL of dNTP mix (10 mM each) L; 1 µL of<br />

sense primer (10-µM stock); 1 µL of antisense primer (10-µM stock); 1 µL of KlenTaq<br />

Advantage; and 30.75 µL of ddH 2 O.<br />

2. Aliquot 40 µL in thin-wall 0.5-mL PCR tubes and keep at room temperature. Add 10 µL<br />

of the DNA template to the mix. Overlay with 40 µL of mineral oil.<br />

3. Cycling parameters. We use a Robocycler 40; during each cycle of PCR we use the<br />

following parameters (see Note 6): denaturation: 99°C × 35 s; annealing: 67°C × 30 s; and<br />

elongation: 68°C × (optimal time for the targeted amplicon; see Note 7).<br />

3.2. Nested Long PCR<br />

1. We have shown that the strategy of nested PCR can be applied with success to long PCR.<br />

It requires a slightly modified master mix, including per reaction (see Note 8): 4.5 µL<br />

of 10× KlenTaq PCR buffer; 1.25 µL of dNTP mix (10 mM each); 1 µL of sense primer<br />

(10 µM); 1 µL of antisense primer (10 µM); 1 µL of KlenTaq Advantage; and 36.25 µL<br />

of ddH 2 O.<br />

2. Aliquot 45 µL of master mix in 0.5-mL thin-wall PCR tubes and overlay with 40 µL of<br />

mineral oil. Add 5 µL of the first-round PCR (see Notes 9 and 10).<br />

3. Cycling parameters as in Subheading 3.1.<br />

3.3. Long RT-PCR<br />

3.3.1. Reverse Transcription<br />

1. Components of the reverse transcription are prepared in a master mix. For each reaction<br />

use (see Note 11) the following: 4 µL of 5× RT buffer; 0.5 µL of RNasin; 1 µL of DTT<br />

(100 mM); 1 µL of dNTP mix (10 mM each); 2.5 µL of primer (10 µM; see Note 12);<br />

and 1 µL of Superscript II.<br />

2. Heat the 10 µL of RNA aliquot at 65°C for 2 min, and then put on ice.<br />

3. Add 10 µL of master mix and incubate 1 h at 42°C (see Note 13).<br />

4. Put on ice, add 1 µL of RNase H and 1 µL of RNase T1, and incubate 20 min at 37°C<br />

(see Note 14).<br />

5. Keep on ice until used in long PCR or keep frozen for later use.<br />

3.3.2. Long RT-PCR<br />

Because of buffer incompatibility (the buffer for Superscript II contains KCl), not<br />

all of the RT reaction can be used in the PCR. We obtain good results by transferring<br />

a small amount of the RT reaction into a long PCR mix. The remainder of the RT<br />

reaction can then be frozen and used at a later time.<br />

1. To amplify by long PCR the cDNA produced in the RT, we prepare a master mix as in<br />

Subheading 3.1. but corrected for adding the template in a volume of 2 µL. For each<br />

reaction, use the following: 5 µL of 10 × KlenTaq PCR buffer; 1.25 µL of dNTP mix<br />

(10 mM each); 1 µL of sense primer (10-µm stock); 1 µL of antisense primer (10-µm<br />

stock); 1 µL of KlenTaq Advantage; 38.75 µL of ddH 2 O.


Long PCR Methodology 175<br />

2. Aliquots (48 µL) of the master mix are placed in thin-wall 0.5-mL PCR tubes and overlaid<br />

with 40 µL of mineral oil.<br />

3. Finally, 2 µL of the RT reaction is added. We neglect the contribution of the 2 µL of RT<br />

reaction to primer and dNTP concentrations, etc.<br />

4. Notes<br />

1. The KlenTaq Advantage polymerase mix also contains anti-Taq antibody, which ensures<br />

a “hot start” to the PCR. We have not found it necessary to include an initial prolonged<br />

denaturation step, so the way we are using the polymerase mix is in fact “time-release<br />

PCR.”<br />

2. The Clontech buffer is Tricine-based, and therefore is not subject to the same drop in pH<br />

at high temperatures that is seen with Tris-based buffers (8). The buffer already contains<br />

Mg 2+ ; we have not found it necessary to modify the concentration of Mg 2+ for any of the<br />

templates we have amplified. The buffer does contain KOAc instead of KCl, the latter<br />

having been shown to decrease the processivity of DNA polymerases (8).<br />

3. Primers: 10 pmol of each primer are used in the PCR. Because long PCR in general uses<br />

a higher annealing temperature than does standard PCR, a requirement of primer design is<br />

that the Tm must be high enough to hybridize at that temperature. Whereas 20 to 25 mers<br />

with a sufficiently high GC content will work, we have also often used primers of 30 to 40<br />

bp without problems, and when incorporating promoters and cloning sites, primers of up<br />

to ~60 bp. Barnes (5) has also reported the use of amplicons of a few hundred bp as “mega<br />

primers.” Primer design should include consideration of secondary structures, dimers, etc.,<br />

but because of the high temperatures throughout the PCR cycle and the “hot start,” there<br />

is in fact greater tolerance for weak secondary structures than with many “standard” PCR<br />

protocols. Because of the high annealing temperature, however, there is a low tolerance<br />

for mismatches between the primers and the template.<br />

4. Long PCRs usually require rapid temperature transition (however, some templates are more<br />

tolerant than others), and several parameters are optimized toward this goal, including the<br />

small reaction volume (50 µL) and the small volume of oil. Because different volumes lead<br />

to different thermal capacity, great consistency is required for optimization. We use the<br />

Stratagene thin-wall tubes; we have found that some other brands may require, for some<br />

templates, different parameters: again, consistency is the essential element.<br />

5. Long PCR requires a thermal cycler with fast temperature transitions; depending on the<br />

template and the size of the target for amplification, not all thermal cyclers will allow<br />

successful amplification (5). In any case, cycling parameters must be determined for each<br />

different model of thermal cycler.<br />

6. We usually use 35 cycles of PCR amplification. If greater yield or sensitivity is required,<br />

consider using nested PCR.<br />

7. The elongation time is usually the only parameter to adjust in the protocol and depends on the<br />

template and the size of the targeted amplicon. Elongation times shorter or longer than the<br />

optimal time can result in sensitivity loss (or even negative results). Extension times that<br />

are too long are often associated with a “smearing” of the reaction product when it is<br />

electrophoresed on agarose gels. Optimal elongation time is best determined by using a<br />

serial dilution of the template (since suboptimal times may give positive reactions but with<br />

a loss of a few logs in sensitivity). The initial time should be chosen between 1 and 2 min<br />

per kb. Like other authors (8), we have found that step-wise increase of the elongation<br />

time can be beneficial. Examples of elongation times for various templates can be found<br />

in Barnes (5) and Tellier et al. (1). For example, to obtain amplicons from the tobacco<br />

mosaic virus (6.2 kb), HAV (7.5 kb), and HCV (9.25 kb), we found we could use the same<br />

elongation times: 9 min 45 s for the first 15 cycles, 11 min for the next 10 cycles, and


176 Tellier et al.<br />

13 min for the last 10 cycles (1). To minimize recombination events during PCR, one<br />

should probably err on the side of longer elongation times.<br />

8. The master mix is merely corrected for a 5 µL volume of template, taking into account the<br />

amount of buffer carried from the first reaction.<br />

9. For nested PCR, the primers of the second round must be internal to those of the first round<br />

(although they can overlap). If the same primers were to be used, artifactual bands produced<br />

by long PCR because of false priming would also get re-amplified. A second round of long<br />

PCR with the same primers can be performed, but this necessitates purification of the first<br />

round amplicon by agarose gel electrophoresis (9).<br />

10. The overall sensitivity that can be achieved can be close to standard PCR: for the lambda<br />

phage DNA, we have obtained an 11-kb amplicon from as few as 200 copies. When we<br />

used a nested long PCR protocol we could obtain, at the end of the whole process, a 5-kb<br />

amplicon starting from as few as 20 copies (1).<br />

11. This protocol assumes that the RNA is diluted in 10 mM DTT and 5% (vol/vol) RNasin<br />

(20–40 U/µL, Promega). If one uses RNA dissolved in RNase free water, increase the DTT<br />

in the RT master mix to 2 µL, and decrease the amount of primer to 1.5 µL.<br />

12. The primer for the reverse transcription must be a template-specific primer and can be one<br />

of the primers used in the long PCR. Do not use random hexamers.<br />

13. The optimal temperature for Superscript II is between 42 and 45°C (10), but it is active<br />

up to 50°C. Although incubation at 50°C was less sensitive in our hands (1), for some<br />

templates with strong secondary structures, it may be advantageous.<br />

14. The treatment with RNase H and RNase T1 is not absolutely necessary, but we have found<br />

that it increases the sensitivity of the long RT-PCR (1). Other authors have also reported<br />

benefits from the use of RNase H (11–13). We have not tested separately the contributions<br />

of RNase H and RNase T1, but the benefits of RNase T1 are expected to vary depending<br />

on the amount of extraneous RNA present in the template.<br />

References<br />

1. Tellier, R., Bukh, J., Emerson, S. U., Miller, R. H., and Purcell, R. H. (1996) Long PCR and<br />

its application to hepatitis viruses: Amplification of hepatitis A, hepatitis B and hepatitis C<br />

virus genomes. J. Clin. Microbiol. 34, 3085–3091. (Erratum published in J. Clin. Microbiol.<br />

1997;35:2713).<br />

2. Duckmanton, L. M., Tellier, R., Liu, P., and Petric, M. (1998) Bovine torovirus: Sequencing<br />

of the structural genes and expression of the nucleocapsid protein of Breda virus. Virus<br />

Res. 58, 83–96.<br />

3. Martino, T. A., Tellier, R., Petric, M., Irwin, D. M., Afshar, A., and Liu P. (1999) The<br />

complete consensus sequence of coxsackievirus B6 and generation of infectious clones by<br />

long RT-PCR. Virus Res. 64, 77–86.<br />

4. Chen, B., Rigat, B., Curry, C., and Mahuran, D. J. (1999) Structure of the GM2A gene:<br />

identification of an exon 2 nonsense mutation and a naturally occurring transcript with an<br />

in-frame deletion of exon 2. Am. J. Hum. Genet. 65, 77–87.<br />

5. Barnes, W. M. (1994) PCR amplification of up to 35-kb DNA with high fidelity and high<br />

yield from λ bacteriophage templates. Proc. Natl. Acad. Sci. USA 91, 2216–2220.<br />

6. Lawyer, F. C., Stoffel, S., Saiki, R. K., Chang, S. Y., Landre, P. A., Abramson, R. D., et al.<br />

(1993) High-level expression, purification, and enzymatic characterization of full-length<br />

Thermus aquaticus DNA polymerase and a truncated form deficient in 5′ to 3′ exonuclease<br />

activity. Genome Res. 2, 275–287.<br />

7. Clontech (1998) User Manual for Advantage cDNA PCR kit and Advantage cDNA<br />

Polymerase Mix.


Long PCR Methodology 177<br />

8. Cheng, S., Fockle, C., Barnes, W. M., and Higuchi, R. (1994) Effective amplification of<br />

long targets from cloned inserts and human genomic DNA. Proc. Natl. Acad Sci. USA<br />

91, 5695–5699.<br />

9. Tellier, R., Bukh, J., Emerson, S. U., and Purcell, R. H. (1997) Amplification of the<br />

full length hepatitis A virus by long RT-PCR and generation of infectious RNA directly<br />

from the amplicon in Viral Hepatitis and Liver Disease, Proceedings of the IX Triennial<br />

International Symposium on Viral Hepatitis and Liver Disease (Rizzetto M., Purcell R. H.,<br />

Gerin J. L., and Verme, G., eds.), Minerva Medica, Turin, pp. 48–50.<br />

10. Gibco BRL (1994) Technical Bulletin 18064-2. Gibco BRL, Gaithersburg, MD.<br />

11. Nathan, M., Mertz, L. M., and Fox, D. K. (1995) Optimizing long RT-PCR. Focus 17,<br />

78–80.<br />

12. Wang, L.-F., Radkowski, M., Vargas, H., Rakela, J., and Laskus, T. (1997) Amplification and<br />

fusion of long fragments of hepatitis C virus genome. J. Virol. Methods 68, 217–223.<br />

13. Lindberg, A. M., Polacek, C., and Johansson, S. (1997) Amplification and cloning of<br />

complete enterovirus genomes by long distance PCR. J. Virol. Methods 65, 191–199.


178 Tellier et al.


Qualitative and Quantitative PCR 181<br />

30<br />

Qualitative and Quantitative PCR<br />

A Technical Overview<br />

David Stirling<br />

1. Introduction<br />

The nature of the polymerase chain reaction (PCR) process lends itself well to<br />

qualitative determinations. It transforms very small quantities of analyte into the realms<br />

of bucket chemistry, allowing specific gene portions to be directly visualized with<br />

ethidium bromide and ultraviolet light. It was these approaches that were first to be<br />

exploited in DNA analysis. The early PCR tests were for the presence or absence of a<br />

gene, transcript, or by inference whole organism (e.g., pathogen identification). These<br />

qualitative tests rapidly evolved to distinguish between related genes or organisms and<br />

accelerated the whole process of molecular taxonomy.<br />

In contrast to this global utility and acceptance, the use of quantitative PCR has<br />

grown more slowly, for similar reasons. A chain reaction by definition is intuitively out<br />

of control. Surely the vagaries of chaos theory will obliterate any hope of quantitation;<br />

it matters not if the butterfly flaps once or one thousand times. Despite these early<br />

misgivings, a great many successful quantitative PCR protocols have been developed<br />

and gained general acceptance.<br />

2. Qualitative PCR<br />

There a number of considerations that should be taken into account when performing<br />

qualitative PCR. Careful thought at the outset of the design process can avoid a great<br />

deal of subsequent frustration. When the PCR product contains the expected amplicon<br />

(assuming all negative controls have been included and are negative), it is safe to<br />

assume the template was present in the starting sample. However, the converse is not<br />

true. There are a number of common causes for the failure of qualitative PCR other<br />

than the absence of template.<br />

2.1. Template Concentration<br />

Even the most basic qualitative PCR is dependent on quantitative changes in template<br />

concentration. The whole process is based on an ideally exponential amplification<br />

of starting material to a point where it can be readily seen and manipulated. Clearly,<br />

this ability will be influenced by both the starting concentration of template and the<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

181


182 Stirling<br />

efficiency of the reaction. Although it is theoretically possible to amplify from a single<br />

copy of a sequence, if the reaction conditions are not optimized, such amplification<br />

may not reach a threshold for detection within the average 30-cycle PCR. If the object<br />

of the exercise is to determine the presence or absence of a specific sequence, elements<br />

of quantitation must be built into the reaction design to determine the required level of<br />

sensitivity. Where PCR is to be used to test for the presence of a pathogen for instance,<br />

consideration has to be given to the clinically significant levels. If a single organism<br />

is clinically significant, then clearly the test has to be capable of detecting a single<br />

copy. If, however, 1000 organisms are required for clinically significant event, the<br />

assay need not be as sensitive (1).<br />

2.2. Mixed Template Competition<br />

If the template contains more than one species of DNA capable of being amplified by<br />

the primers, those templates that exist in greater initial concentrations or that amplify<br />

more efficiently may outcompete the remaining templates. For instance, PCR is used to<br />

amplify papilloma virus sequences in cervical cytology specimens. Consensus primers<br />

are used to amplify a broad spectrum of viral strains, which can then be characterized<br />

by restriction digest or probe hybridization. If the patient has a mixed infection, it<br />

is likely that the more abundant strain will be amplified preferentially, resulting in<br />

only one strain being represented in the PCR product (2). This is perfectly adequate to<br />

determine which are the predominant strain(s) within a sample. However, if a complete<br />

description of the strains present is required, PCR should be performed to specifically<br />

amplify individual strains<br />

3. Quantitative PCR: General Considerations<br />

3.1. End-Point Analysis<br />

Unfortunately, the cause of quantitative PCR has not promoted by the use of end-point<br />

analysis. This approach simply amplifies multiple templates under the same set of conditions<br />

and examines the amount of product at the end of the run. Because PCR exhibits a<br />

typical exponential amplification of product, complete with lag phase and plateau, it is<br />

easily possible for widely differing starting template concentrations to yield remarkably<br />

similar final product concentrations. This can be a useful technique but only if great<br />

care is taken to establish an appropriate end point. Reactions should be sampled after<br />

a wide range of cycle numbers, to delineate the exponential phase, and the subsequent<br />

plateau. Tests should then be designed to sample during the exponential phase of<br />

amplification, before primers, dNTPs, and or enzyme become limiting. As with all<br />

quantifications, it is greatly enhanced by the inclusion of a suitable internal control.<br />

3.2. Limiting Dilution<br />

PCR quantitation can be achieved using qualitative end points. For the PCR to<br />

proceed, there must me a theoretical minimum of one template per reaction. Thus,<br />

if a dilution series is performed on the template, a dilution can be reached where an<br />

aliquot of sample taken for PCR has a statistical probability of containing no template<br />

molecules. The higher the initial template concentration, the greater the dilution<br />

required needed to reach that point. This approach has been used with great success<br />

to quantify viral titers (3).


Qualitative and Quantitative PCR 183<br />

3.3. Competitive Template PCR<br />

If two different templates capable of being amplified by the same primers are<br />

present at the start of a PCR, one will tend to predominate by the end. This will be<br />

determined by a number of factors, such as relative starting concentrations, size, and<br />

internal sequence. If a competitive template is designed to be almost identical to the<br />

test sequence, the final dominance of test or competitor will entirely be the result of the<br />

initial concentration. Many groups, including myself, have used competitors identical<br />

to the test DNA apart from one base change (used to introduce a restriction site to<br />

distinguish the templates) in this way (4). A dilution series of competitor is set up with<br />

a constant concentration of test sample and subjected to PCR. A plot of the relative<br />

product concentrations at the end of the amplification reveals a point where there is<br />

equivalent production from both template and competitor. This then yields the concentration<br />

of the test sample template. This is a very robust approach to quantitation,<br />

but is time consuming both to establish and to run, requiring many PCRs for each<br />

sample, and being difficult to automate.<br />

3.4. Real-Time PCR<br />

The development of fluorescent detection systems, capable of monitoring PCR<br />

product accumulation while it occurs has greatly improved the reliability of quantitative<br />

PCR. There is a well-established inverse correlation between the template concentration<br />

and the duration of the lag phase of the PCR. An entire generation of automated systems<br />

has been developed to exploit this. Using any of several different chemistries, the<br />

accumulation of PCR product is monitored until a predetermined threshold is reached.<br />

This lag time is then compared with a standard curve and a concentration calculated.<br />

Lending itself very readily to automation, real-time PCR is becoming the method<br />

of choice for most quantitative PCR systems. Although the equipment for performing<br />

such analysis has been fairly expensive, second-generation systems are already on the<br />

market at much more reasonable costs.<br />

One final note of caution: It is not uncommon for these systems to be based upon the<br />

detection of accumulated PCR product by the use of a fluorescent dye, such as SYBR<br />

green. Such systems can work very well when the PCRs are optimized to yield only<br />

the product to be quantified. If, however, a great many heterogeneous samples are to be<br />

analyzed, it is entirely possible that some will yield additional bands, confounding the<br />

results. The use of hybridization probe strategies will greatly reduce this possibility.<br />

References<br />

1. Tonjum, T., Klintz, L., Bergan, T., Baann, J., Furuberg, G., Cristea, M., et al. (1996)<br />

Direct detection of Mycobacterium tuberculosis in respiratory samples from patients in<br />

Scandinavia by polymerase chain reaction. Clin. Microbiol. Infect. 2, 127–131.<br />

2. O’Leary, J. J., Landers, R. J., Crowley. M., Healy, I., O’Donovan, M., Healy, V., et al.<br />

(1998) Human papillomavirus and mixed epithelial tumors of the endometrium. Hum.<br />

Pathol. 29, 383–389.<br />

3. Lee, T. H., Sunzeri, F. J., Tobler, L. H., Williams, B. G., Busch, M. P. Lee, T. H., et al.<br />

(1991) Quantitative assessment of HIV-1 DNA load by coamplification of HIV-1 gag and<br />

HLA-DQ-alpha genes. AIDS 5, 683–691.<br />

4. Stirling, D., Hannant, W. A., and Ludlam, C. A. (1998)Transcriptional activation of the<br />

factor VIII gene in liver cell lines by interleukin-6. Thromb. Haemost. 79, 74–78.


184 Stirling


PCR Detection of Tumor Cells 185<br />

31<br />

Ultrasensitive PCR Detection of Tumor Cells in Myeloma<br />

Friedrich W. Cremer and Marion Moos<br />

1. Introduction<br />

Chromosomal aberrations, such as translocations or inversions, described for a<br />

growing number of malignancies, are now widely used to detect tumor cells by<br />

polymerase chain reaction (PCR). However, in multiple myeloma (MM), no such<br />

ubiquitous PCR marker exists. Therefore, other means have been established to<br />

distinguish myeloma cells from normal cells. Because the plasma cells of a myeloma<br />

clone share an identical rearranged immunoglobulin gene sequence, it is possible to<br />

detect malignant cells with PCR primers specific for the VDJ rearrangement of the<br />

heavy chain of each myeloma clone. The sensitivity and specificity of this method,<br />

named allele-specific oligonucleotide (ASO) PCR, even with low proportions of<br />

malignant cells, has been proven (1).<br />

The heavy chains of the immunoglobulins have three highly variable regions near<br />

their amino-terminal end that mediate specific binding of the antigen. These regions<br />

are called complementarity determining regions (CDR1 to –3). The sequences of the<br />

heavy chain of the immunoglobulin genes are encoded on chromosome 14. In germline<br />

configuration, several hundred base pairs divide the approx 200 variable regions<br />

(V) from the 30 diversity regions (D). Several kilobase pairs downstream, 6 joining<br />

regions (J) are located. About 7000 base pairs to the 3′-end the constant regions of the<br />

heavy chain are encoded. During B-cell maturation, a random rearrangement of<br />

these regions occurs, in which one of the V-, one of the D-, and one of the J-regions are<br />

joined together, thus generating the VDJ-segment. Together with the sequence of the<br />

rearranged light chain, it determines the antigenic specificity of the immunoglobulin.<br />

The rearrangement of V-, D-, and J-regions can generate over 35,000 different VDJsequences.<br />

The diversity of these sequences is further increased by three different<br />

mechanisms: (1) the recombination process is imprecise, thus nucleotides can be lost<br />

or stretches of bases that divide the different regions are not deleted; (2) nucleotides<br />

can be inserted without matrix at the junctions of V-, D-, and J-regions; and (3) somatic<br />

hypermutations, especially of the V-regions, further enhance the antigenic specificity of<br />

the immunoglobulin. Although CDR1- and CDR2-regions are entirely encoded by the<br />

V-region, the CDR3-region stretches over the 3′-end of the V-region, the D-region<br />

and the 5′-end of the J-region (see Fig. 1). Thus, the CDR3-region has the highest<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

185


186 Cremer and Moos<br />

Fig. 1. Rearranged VDJ segment of a mature B-cell and strategies for consensus PCR to<br />

amplify the VDJ segment. The CDR regions of the heavy chain of the immunoglobulin gene<br />

are flanked by highly conserved segments, the framework (FR) regions 1 through 4. Consensus<br />

primers complementary to these FR-regions have been developed that can be used to amplify<br />

the enclosed highly variable CDR regions regardless of their sequence. In this chapter, two<br />

strategies are described using either FR1C or FR3A plus LJH.<br />

diversity of all CDR-regions, and its sequence is a specific marker for each clone<br />

of B-cells (2,3).<br />

In MM, the CDR regions of the malignant cells are somatically hypermutated, and<br />

no further mutations occur, which would lead to an oligoclonal diversification (4). This<br />

is a prerequisite for designing ASO primers complementary to the CDR regions of the<br />

malignant clone that allow the detection of myeloma cells by PCR.<br />

For designing ASO primers, the sequence of the CDR regions of the malignant<br />

clone has to be determined. In a first step, the CDR regions of virtually all B-cells are<br />

amplified with consensus primers flanking the CDRs (see Fig. 1). Besides the strategies<br />

using FR1C (5) or FR3A (6) as sense primers and LJH as antisense primer, consensus<br />

PCR with family-specific primers complementary to the leader segment 5′-end to<br />

the V-region (7), with a mixture of six FR1 family-specific primers (8), with an FR2<br />

consensus primer alone (9) or as a mixture with the FR1 family-specific primers<br />

VH5 and VH6 (5) and a mixture of JH1245, JH3, and JH6 antisense primers (5)<br />

has been described. After consensus PCR, the product of the malignant clone can<br />

be distinguished from normal ones clones by its predominant occurrence among<br />

the polyclonal CDR-regions. After sequencing (directly or after cloning), primers<br />

complementary to the CDR regions can be designed. These ASO primers have to be<br />

tested for specificity and sensitivity before they can be used for the PCR detection of<br />

cells of the malignant clone. PCR is quantified using the method of limiting dilutions.<br />

2. Materials<br />

2.1. Isolation of Nucleic Acids<br />

1. Bone marrow (BM) sample with a high proportion of myeloma cells.<br />

2. Ficoll separating solution (<strong>Bio</strong>chrom, Berlin, Germany).<br />

3. Phosphate-buffered saline (pH 7.4; Gibco BRL, Eggenstein, Germany).<br />

4. TE buffer (10 mM Tris-HCl, pH 7.5, and 1 mM EDTA).<br />

5. Reagents for RNA extraction, for example, Trizol Reagent (Gibco BRL).<br />

6. Reagents for DNA extraction, for example, DNAzol (Gibco BRL).<br />

7. Ethidium bromide-stained agarose gels (0.8 and 2%).


PCR Detection of Tumor Cells 187<br />

8. Denaturating agarose gels (1.2%) containing 2.2 mM formaldehyde for RNA.<br />

9. PCR reagents and primers for amplifying reference genes like β-actin or GAPDH.<br />

2.2. Consensus RT-PCR<br />

1. GeneAmp RNA PCR Core Kit (Perkin–Elmer, Weiterstadt, Germany).<br />

2. Amplitaq Gold DNA polymerase with GeneAmp 10× PCR buffer II (Perkin–Elmer) and<br />

MgCl 2 solution.<br />

3. Mixture of dATP, dCTP, dGTP, and dTTP, concentration 2.5 mM each (stock solution<br />

100 mM, Promega, Madison WI).<br />

4. Consensus primers FR1C (5′-GGTGCAGCTGS(A/T)GSAGTC(A/G/T)GG-3′) (5), FR3A<br />

(5′-ACACGGCYSTGTATTACTGT-3′) (6), and LJH (5′-TGAGGAGACGGTGACC-3′) (6).<br />

5. Ethidium bromide-stained agarose gels (2 and 5%, NuSieve 3:1 Agarose, FMC/<strong>Bio</strong>zym,<br />

Oldendorf, Germany).<br />

2.3. Identification of the Myeloma CDR Regions<br />

1. Scalpel.<br />

2. DNA purification kit (EasyPure DNA Purification kit, <strong>Bio</strong>zym, Göttingen, Germany).<br />

3. Cloning kit for PCR products (TOPO TA Cloning Kit, Invitrogen, De Schelp, The<br />

Netherlands).<br />

4. LB medium, LB plates, with 50 µg/mL kanamycin (Boehringer Mannheim, Mannheim,<br />

Germany).<br />

5. Lysis buffer (100 µg/mL of Proteinase K (Boehringer Mannheim) in 10 mM Tris-Cl,<br />

pH 7.5, and 1 mM EDTA).<br />

6. PCR reagents (see Subheading 2.2.) and M13 universal (forward) and M13 reverse<br />

primers.<br />

7. Sequitherm Cycle Sequencing kit (<strong>Bio</strong>zym, Göttingen, Germany).<br />

2.4. Quantitative PCR (qPCR) Using ASO Primers<br />

1. DNA processing software, for example Gene Jockey II (<strong>Bio</strong>soft, Cambridge, United<br />

Kingdom).<br />

2. PCR reagents, ASO primers as designed, LJH primer, ethidium bromide stained agarose<br />

gels (2% or 5%).<br />

3. Buffy coat DNA from healthy donors.<br />

4. Computer program for likelihood maximization and χ 2 minimization, for example,<br />

MAXLIKE. This program written in C (Watcom C/C++ version 10.6, Powersoft, Waterloo,<br />

Canada) running under DOS on an IBM compatible PC or the source code can be<br />

obtained for free according to the terms and conditions for copying, distribution, and<br />

modification of programs under the GNU general public license agreement from the<br />

authors (see contact address).<br />

5. Myeloma or B-cell lines, for example U266 (10) or Daudi (11).<br />

3. Methods<br />

3.1. Isolation of Nucleic Acids<br />

1. Collect bone marrow aspirate in two 10-mL syringes with heparin added for anticoagulation.<br />

This will be the starting material for both the identification of the VDJ sequence of<br />

the malignant clone and the testing of ASO primers.<br />

2. Isolate mononucleated cells by centrifugation over Ficoll separating solution.<br />

3. Wash cells twice in phosphate-buffered saline, pelleting between washes by centrifugation<br />

at 800g for 10 min.


188 Cremer and Moos<br />

4. Count cells and divide them into at least 1 × 10 7 cells for RNA extraction and the remaining<br />

cells for DNA extraction. If the cell pellet appears red, resuspend in TE buffer and vortex<br />

for 5 to 10 s to lyse remaining red blood cells, then pellet again quickly and remove<br />

TE buffer.<br />

5. Add 250 µL of Trizol reagent per 1 × 10 7 cells then freeze cells at –20°C for later RNA<br />

extraction. Freeze the remaining cells immediately for later DNA extraction.<br />

6. Thaw pelleted cells in Trizol reagent and extract RNA according to the manufacturers’<br />

instructions.<br />

7. Thaw pelleted cells and isolate DNA, for example using the DNAzol reagent according to<br />

the manufacturers instructions or see Chapter 6 for DNA extraction protocol.<br />

8. Determine concentration of RNA and DNA by optical density (OD) measurement at 260<br />

and 280 nm. Adjust concentration of RNA to 500 ng/µL and of DNA to 100 ng/µL.<br />

9. Confirm integrity of RNA by electrophoresis of 500 ng on a denaturing 1.2% agarose gel<br />

containing 2.2 mM formaldehyde.<br />

10. Confirm integrity of DNA by electrophoresis of 100 ng on a 0.8% agarose gel.<br />

11. Check quality of DNA or RNA by PCR amplifying a gene like β-actin in the case of DNA<br />

or a housekeeping gene like GAPDH in the case of RNA.<br />

3.2. Consensus RT-PCR<br />

1. Perform reverse transcription reaction in a total volume of 20 µL containing 2 µg of RNA<br />

using commercially available kits.<br />

2. Prepare a PCR mixture containing 8 µL of 10× PCR buffer, 1 µL of 20 mM primer<br />

solutions FR1C or FR3A plus LJH-CL, and 2.5 U Amplitaq Gold DNA polymerase.<br />

3. Add the total volume of the RT reaction mixture to the PCR mixture. The final volume<br />

should be 100 µL.<br />

4. Amplify with a program consisting of 7 min of denaturation and enzyme activation at<br />

94°C, followed by 40 cycles of denaturation for 1 min at 94°C and combined annealing<br />

and extension at 63°C (for primer FR3A) or 65°C (for primer FR1C) for 1 min, followed<br />

by a final extension step at 65°C for 5 min.<br />

5. Resolve PCR products either on a 2% (PCR products are approx 350 bp in size if FR1C<br />

was used) or 5% (PCR products are approx 110 bp in size if FR3A was used) ethidium<br />

bromide stained agarose gels. The monoclonal CDR regions lead to a distinct band if the<br />

proportion of myeloma cells is high enough. This band can be distinguished from the<br />

surrounding smear of polyclonal CDR-regions. Figure 2 shows an example of a consensus<br />

PCR (see Note 1).<br />

3.3. Identification of the Myeloma CDR Regions<br />

3.3.1. Cloning of Consensus PCR Products and Direct Lysis<br />

of Transformed Bacteria<br />

1. Excise appropriately sized consensus PCR products from the agarose gel with a scalpel (4).<br />

2. Purify DNA from the agarose block using commercially available kits.<br />

3. Clone-purify PCR products into plasmids and transform competent cells using kits<br />

featuring TA-cloning. If possible, use kanamycin rather than ampicillin for selection, as<br />

satellite colonies occur less frequently. Grow bacteria for 12 to 16 h at 37°C.<br />

4. Pick single colonies and transfer to a replica plate. Spread on an area of approx 1 cm 2 .<br />

Incubate for 12 to 16 h at 37°C. If a distinct band without surrounding smear was visible<br />

after consensus PCR, 10 to 20 colonies should be sufficient. If a band surrounded by a<br />

polyclonal smear was visible, 20 to 40 colonies should be picked.


PCR Detection of Tumor Cells 189<br />

Fig. 2. Consensus PCR using primers FR3A and LJH. In 5 of 16 cases, the consensus PCR<br />

product of the malignant clone can be distinguished from the surrounding polyclonal smear.<br />

In all other cases, only a smear, like the one seen in buffy coat DNA, is visible. This figure<br />

also illustrates the low rate of distinct consensus PCR products obtained if using DNA instead<br />

of RNA. Detection rates can further be increased by using two different consensus strategies<br />

in parallel.<br />

5. Scrape layer of bacteria from the plate and suspend in 100 µL of lysis buffer. Vortex<br />

vigorously.<br />

6. Incubate the suspension at 65°C for 15 min.<br />

7. Inactivate proteinase K at 95°C for 15 min.<br />

8. Pellet debris by centrifugation at maximum speed and 4°C. Transfer the supernatant<br />

containing the plasmids to a fresh tube. Store at 4°C.<br />

9. Analyze plasmids for inserts by PCR using primers complementary to the M13 sites<br />

of the vector. Set up a PCR containing 5 µL of 10× PCR buffer, 2 mM MgCl 2 , each<br />

deoxynucleotide at 0.1 mM, primers M13 forward plus M13 reverse at 0.4 µM each, 2.5 U<br />

Taq-DNA-polymerase, and 1 µL of plasmid solution. Amplify for 40 cycles of 1-min<br />

denaturation at 94°C, 1 minute annealing at 65°C, and extension at 72°C, followed by a<br />

final extension of 5 min at 72°C. Analyze on a 2% agarose gel.<br />

3.3.2. Sequencing and Identification of the CDR Regions of the Myeloma Clone<br />

1. Prepare sequencing reactions using commercially available kits according to the manufacturer’s<br />

instructions with 4 to 8 µL of plasmid solution, and perform cycle sequencing.<br />

2. Analyze on a sequencer, for example Alf-express (Pharmacia, Freiburg, Germany).<br />

3. Analyze obtained sequences by aligning them to known CDR regions (examples for CDR3<br />

regions are given in Fig. 3; see Note 2).<br />

4. Identify the consensus PCR product of the malignant clone by its predominant occurrence<br />

among the polyclonal CDR-regions. B-cells of normal clones are not encountered more<br />

frequently than once in 20000 B-cells (12). At least five identical clones should be typed<br />

to identify the myeloma clone.


190 Cremer and Moos<br />

Fig. 3. Alignment of sequences of CDR3 consensus PCR products of the malignant clones<br />

from 10 patients with MM and an example of four ASO primers designed for the last of the<br />

given sequences (testing of the primers is shown in Fig. 4). Aligning allows to distinguish<br />

variable and rather conserved segments. Primers are designed to be complementary to the<br />

highly variable region.<br />

3.4. Quantitative PCR (qPCR) Using ASO Primers<br />

To quantitate PCR results, essentially four different methods have been described.<br />

Although the measurement of the amount of PCR product, the coamplification of a<br />

control gene, and the competitive PCR rely on the quantitation of the generated product,<br />

the limiting dilution assay analyzes only positive and negative PCRs at different<br />

dilution levels (13). Amplification efficiencies and hence the amounts of PCR product<br />

can vary substantially from tube to tube despite identical and simultaneous processing<br />

of reaction mixtures (14), even if quantitation is performed in the exponential phase of<br />

amplification. Limiting dilutions are less prone to errors caused by these deviations.<br />

Standards can be used to control for these differences; however, in MM a system of<br />

standard and template would have to be established for each new patient. Therefore,<br />

we prefer the limiting dilutions assay, which analyzes only PCR positivity versus<br />

PCR negativity. A prerequisite of this method is that a single copy of target DNA<br />

can be detected.<br />

3.4.1. Designing Primers (see Note 3)<br />

1. ASO primers should be 18 to 23 bases long.<br />

2. The initiation of the Taq-DNA-polymerase can be enhanced by 3′-ends with NS or SS as<br />

bases, but more than 2 G at the 3′-end should be avoided.<br />

3. The 5′-end and 3′-end of primers should not be complementary to each other to avoid<br />

primer homodimerization.<br />

4. The A/T to G/C ratio should be approx 50%, if possible.<br />

5. There are several computer programs that can be used to check primers for hairpin<br />

formation and primer pairs for dimerization.<br />

6. FR3-strategy: Identify hypervariable parts of the CDR3-region of the malignant clone<br />

by aligning the sequence to known CDR3-sequences. Avoid segments that are relatively<br />

conserved. Design ASO primer as forward primer according to above considerations. Use<br />

ASO primer together with LJH as antisense primer.


PCR Detection of Tumor Cells 191<br />

7. FR1-strategy: Distinguish FR regions and CDR regions by aligning sequence to known<br />

CDR sequences. Design the first ASO primer as the forward primer complementary to the<br />

CDR1 or CDR2 region (CDR1 regions are often rather conserved, thus CDR2-specific<br />

primers might be preferred). Design the second ASO primer as the antisense primer<br />

complementary to the CDR3 region. Use these primers as a pair for PCR.<br />

3.4.2. Testing of ASO Primers<br />

Because even carefully designed primers may not work as well as expected, only<br />

testing can distinguish between oligonucleotides that fulfill all requirements and<br />

those that are not specific or sensitive enough. CDR1/2- plus CDR3-specific ASO<br />

primer pairs do not necessarily give better results than CDR3-specific plus J-consensus<br />

primers. An example of a primer test is given in Fig. 4. Design several CDR3-specific<br />

or CDR1/CDR2-specific plus CDR3-specific (antisense) primers. An example for 4<br />

ASO primers is given in Fig. 3.<br />

1. Prepare PCRs containing 5 µL of 10× PCR buffer, 2 mM MgCl 2 , each deoxynucleotide at<br />

0.1 mM, the ASO primer pair or the CDR3-specific primer plus LJH-CL at 0.8 µM each,<br />

and 2.5 U Taq-DNA-polymerase.<br />

2. Add 500 ng of DNA from the initial BM sample to generate a positive control. Add 1000 ng<br />

of buffy coat DNA from healthy donors to generate a negative control that tests for<br />

specificity of the primers. Adjust with distilled water to a final volume of 50 µL. Repeat<br />

this for every primer combination to be tested.<br />

3. Amplify with a program consisting of 7 min of preheating at 94°C, 60 cycles of 1 min<br />

of denaturation at 94°C, and 1 minute of annealing and extension at 63°C, ending with<br />

a final extension at 63°C for 5 min.<br />

4. Analyze PCR products on a 2 (for FR1/2-strategy) or 5% (for FR3-strategy) ethidium<br />

bromide-stained agarose gel. Reactions containing DNA from the positive control should<br />

display a prominent band of the expected size (see Note 4). No such product should be<br />

visible in those with buffy coat DNA only. Avoid primers that generate a multitude of<br />

nonspecific products.<br />

3.5. Quantitation by Limiting Dilutions<br />

1. Isolate DNA from the sample to be assessed (see Note 3).<br />

2. Determine the DNA concentration of the sample by OD measurement. Add distilled water<br />

to achieve a DNA concentration of 100 ng/µL.<br />

3. Prepare a dilution series in 0.5 log steps (resulting in dilution levels of 13, 110,<br />

133, etc.).<br />

4. For each dilution level, set up a PCR containing 10 µL of the DNA solution. Use conditions<br />

as determined to be optimal for the ASO primers used (see Subheading 3.7.) and<br />

amplify.<br />

5. Analyze PCR products on agarose gels. Determine the highest dilution that still generates<br />

a specific PCR product.<br />

6. Set up five replicates of PCRs containing 10 µL of DNA solution from the highest dilution<br />

level that was PCR positive, five replicates with DNA from the lowest dilution level that<br />

was PCR negative, and five replicates with DNA from the next lowest dilution level that<br />

was PCR negative and amplify.<br />

7. Analyze on an agarose gel.<br />

8. Decrease or increase the dilution level analyzed in five replicates until the pattern of<br />

PCR results finally obtained shows a change from PCR positivity in all replicates (or the


192 Cremer and Moos<br />

Fig. 4. Testing of primers (see Fig. 3) as described in Subheading 3.4.2. For each ASO<br />

primer, a negative control without added DNA, a positive control with DNA from the initial BM<br />

sample, and a specificity test with buffy coat DNA were amplified by PCR. Using primer 1, only<br />

the positive control leads to the PCR product of the expected size. Primers 2, 3, and 4 produce<br />

several nonspecific bands, even in the negative control. Primer 3 generates a false-positive PCR<br />

product of the same size with buffy coat DNA. Primer 1 was selected, and PCR conditions were<br />

further optimized for this oligonucleotide as described in Subheading 3.7.<br />

undiluted level has been reached) to partial PCR positivity in the next higher dilution levels<br />

to PCR negativity in all replicates in the highest dilution levels (see Fig. 5).<br />

9. To deduce the proportion of malignant cells in the undiluted sample from the pattern<br />

of PCR results, analyze by likelihood maximization and by χ 2 -minimization (15), for<br />

example, by using the MAXLIKE computer program. Ten microliters of the undiluted<br />

DNA solution (100 ng/µL) are equivalent to 165,000 cells, which is used as a starting<br />

point for the calculations. The program will return the most probable value for the initial<br />

proportion of malignant cells in the analyzed sample.<br />

10. Compare the values obtained by likelihood maximization and by χ 2 -minimization. If PCR<br />

results are plausible, both methods will yield consistent values (see Note 5).<br />

3.6. Testing the Sensitivity of ASO Primers<br />

A prerequisite for the analysis of the pattern of PCR results by χ 2 -minimization<br />

or likelihood maximization is that a single copy of target DNA per PCR tube can be<br />

detected. It is advisable that at least the overall sensitivity of the assay should be tested,<br />

as described in Subheading 3.6.1.


PCR Detection of Tumor Cells 193<br />

Fig. 5. Example for a qPCR of a PB sample from a patient with MM. The specific PCR<br />

product is 85 bp. The highest dilution level at which all five replicates are PCR positive is<br />

11000, the lowest dilution level at which all five replicates are PCR negative is 110,000.<br />

Analysis of this pattern of PCR results by likelihood maximization and χ 2 -minimization yielded<br />

a result of 1.2% of malignant cells.<br />

3.6.1. Testing the Sensitivity of the Assay Using a Cell Line<br />

1. Design primers for a B-cell or myeloma cell line and test for specificity.<br />

2. Isolate DNA from cells of the cell line and from buffy coat of healthy donors and quantitate<br />

by OD measurement.<br />

3. Mix DNA from the cell line with buffy coat DNA to simulate samples with different<br />

proportions of malignant cells (100%, 10%, 1%, 0.1%, 0.01%, and 0.001%).<br />

4. Analyze all samples by qPCR (see Chapter 6.3.) using the appropriate ASO primers.<br />

5. Compare the results of the qPCR with the simulated tumor loads.<br />

3.6.2. Testing Primer Sensitivity by Comparison to Results of Flow Cytometry<br />

1. Collect a BM sample with a proportion of malignant cells of greater than 2% from a patient<br />

in whom ASO primers have been devised.<br />

2. Assess sample for CD38++, κ/γ-restricted cells by flow cytometry. This gives an approximate<br />

value for the proportion of malignant cells.<br />

3. Isolate DNA from this sample (see Subheading 3.1.).<br />

4. Quantitate the tumor load by qPCR as described in Subheading 3.1.


194 Cremer and Moos<br />

5. Compare the value determined by qPCR with the result of the analysis by flow cytometry.<br />

If the result obtained by qPCR is definitely lower than the one obtained by flow cytometry,<br />

then the ASO primer is most probably not able to detect a single copy of the template<br />

of the malignant clone.<br />

3.7. Optimization of PCR Conditions for ASO Primers<br />

1. As a starting point, use PCRs of 50 µL containing 5 µL of 10× PCR buffer, 2 mM MgCl 2 ,<br />

each deoxynucleotide at 0.1 mM, the ASO primer pair or the CDR3-specific primer<br />

plus LJH at 0.8 µM each, 2.5 U Taq-DNA-polymerase, and a maximum of 1000 ng<br />

of DNA. Amplify with a program consisting of 7 min preheating at 94°C, 60 cycles of<br />

1 min of denaturation at 94°C and 1 min of annealing and extension at 63°C, ending with<br />

a final extension at 63°C for 5 min.<br />

2. If these amplification conditions lead to unsatisfactory results, for example, too many<br />

nonspecific products or low intensity of the specific product, optimize the reaction<br />

conditions. Always amplify a negative control without added DNA, a negative control<br />

with buffy coat DNA from healthy donors, and a positive control with DNA from a BM<br />

sample of the patient.<br />

3. Vary the annealing temperature in steps of 1°C. Vary the MgCl 2 concentration in steps of<br />

0.25 mM. These parameters work synergistically.<br />

4. Vary the concentration of primers in 0.2-µM steps in a range of 0.4 to 1.2 µM. Higher<br />

concentrations of primers can lead to a higher sensitivity; however, this is hampered by<br />

the occurrence of more nonspecific products.<br />

5. Increase the number of amplification cycles to improve sensitivity. Decrease the number<br />

of cycles to reduce nonspecific products. At least 50 cycles should be performed to allow<br />

detection of single copies of template, which is a prerequisite for the statistical analysis of<br />

limiting dilution series by likelihood maximization.<br />

4. Notes<br />

1. No distinct band or multiple bands are visible after consensus PCR: Some BM samples will<br />

not lead to a single distinct band of the expected size after consensus PCR. Either only a<br />

polyclonal smear is visible or, in rare cases, more than one band of the expected size. (1)<br />

Change the strategy used for consensus PCR (FR3 or FR1). Try DNA instead of RNA.<br />

Prepare a PCR mixture containing 10 µL of 10× PCR buffer, 2 mM MgCl 2 , each deoxynucleotide<br />

at 0.05 mM, primers FR1C or FR3A plus LJH at 0.4 µM each, and 2.5 U Taq-<br />

DNA-polymerase. Add 500 to 1000 ng of DNA and distilled water to a final volume of 100 µL.<br />

Amplify with a program consisting of 7 min denaturation and enzyme activation at 94°C,<br />

followed by 50 cycles of denaturation for 1 min at 94°C and combined annealing and<br />

extension at 63°C (for primer FR3A) or 65°C (for primer FR1C), followed by a final<br />

extension step at 65°C for 5 min. If possible, collect another BM aspirate and start again.<br />

2. Sequence has no resemblance to known CDR regions: This is most probably caused by a<br />

nonspecific product of the consensus PCR. Perform consensus RT-PCR again.<br />

3. Sequence of a CDR-region is unsuitable for primer design: Sometimes the sequence of<br />

a CDR region is too short, has too many base repeats, or has too many long stretches<br />

rich in G and C, which makes designing primers difficult. Because CDR1 and CDR2<br />

region are shorter and less variable than the CDR3-region, one should never do without a<br />

CDR3-specific primer. If segments suitable for primer design are not long enough, identify<br />

at least a short segment that is apt. Devise primers complementary to this stretch, and<br />

elongate in 5′-end direction (never in 3′-direction, because this end is more important for<br />

specific annealing of the primer to the template). Use the CDR1/2- plus CDR3-specific<br />

primer strategy, if the CDR3 region is the one that is difficult for primer design.


PCR Detection of Tumor Cells 195<br />

4. The ASO primer does not generate specific PCR product or is not specific or sensitive<br />

enough: Optimize PCR conditions and test primer for specificity or sensitivity. Devise<br />

new primers. Change primer strategy (CDR3 or CDR1/ plus CDR3). If no specific product<br />

can be generated at all, it is most probable that the sequence of the CDR-regions is not<br />

that of the malignant clone.<br />

5. qPCR results are implausible: This is most probably caused by a faulty dilution series.<br />

Prepare dilution series anew and reanalyze the sample by qPCR.<br />

5. Discussion<br />

The described qPCR assay with ASO primers complementary to CDR regions of<br />

the malignant clone can be used to quantitate the proportion of malignant cells in<br />

peripheral blood (PB) and BM, aliquots of leukapheresis products, sorted cell fractions,<br />

or in cultured cells. It is possible to use this method for quantitative follow up in the<br />

course of therapy (16), to determine the tumor load of different stem cell sources (17),<br />

to determine the effect of different mobilization regimens and the duration of collection<br />

on the number of malignant cells in leukapheresis products (18,19) or to assess the<br />

extent of involvement in the malignant process of different cell fractions, for example,<br />

in the CD19+ or CD20+ cells of PB (20,21). The qPCR assay has evolved into an<br />

accurate and very sensitive tool for the detection of malignant cells in MM.<br />

References<br />

1. Cremer, F. W., Kiel, K., Wallmeier, M., Goldschmidt, H., and Moos, M. (1997) A quantitative<br />

PCR assay for the detection of low amounts of malignant cells in multiple myeloma.<br />

Ann. Oncol. 8, 633–636.<br />

2. Tonegawa, S. (1983) Somatic generation of antibody diversity. Nature 302, 575–581.<br />

3. Walter, M. A., Surti, U., Hofker, M. H., and Cox, D. W. (1990) The physical organisation<br />

of the human immunoglobulin heavy chain gene complex. Eur. Mol. <strong>Bio</strong>l. Org. J. 9,<br />

3303–3313.<br />

4. Bakkus, M. H., Heirman, C., Van Riet, I., Van Camp, B., and Thielemans, K. (1992)<br />

Evidence that multiple myeloma Ig heavy chain VDJ genes contain somatic mutations but<br />

show no intraclonal variation. Blood 80, 2326–2335.<br />

5. Aubin, J., Davi, F., Nguyen-Salomon, F., Leboeuf, D., Debert, C., Taher, M., et al. (1995)<br />

Description of a novel FR1 IgH PCR strategy and its comparison with three other strategies<br />

for the detection of clonality in B cell malignancies. Leukemia 9, 471– 479.<br />

6. Brisco, M. J., Tan, L. W., Orsborn, A. M., and Morley, A. A. (1990) Development of a<br />

highly sensitive assay, based on the polymerase chain reaction, for rare B-lymphocyte<br />

clones in a polyclonal population. Br. J. Haematol. 75, 163–167.<br />

7. Campbell, M. J., Zelenetz, A. D., Levy, S., Levy, R. (1992) Use of family specific leader<br />

region primers for PCR amplification of the human heavy chain variable region gene<br />

repertoire. Mol. Immunol. 29, 193–203.<br />

8. Deane, M., McCarthy, K. P., Wiedemann, I. M., and Norton, J. D. (1991) An improved<br />

method for detection of B-lymphoid clonality by polymerase chain reaction. Br. J.<br />

Haematol. 5, 726–730.<br />

9. Diss, T. C., Peng, H., Wotherspoon, A. C., Isaacson, P. G., and Pan, L. (1993) Detection<br />

of monoclonality in low-grade β-cell lymphoma using the polymerase chain reaction is<br />

dependent on primer selection and lymphoma type. J. Pathol. 169, 291–295.<br />

10. Moore, G. E. and Kitamura, H. (1968) Cell line derived from patient with myeloma. N. Y.<br />

State J. Med. 68, 2054–2060.


196 Cremer and Moos<br />

11. Klein, E., Klein, G., Nadkarmi, J. F., Nadkarmi, J. J., Vigzell, H., and Clifford, P. (1968)<br />

Surface IgM-k specificity on a Burkitt lymphoma cell in vivo and in derived culture lines.<br />

Cancer Res. 28, 1300–1310.<br />

12. Yamada, M., Wasserman, R., Reichard, B. A., Shane, S., Caton, A. J., and Rovera, G.<br />

(1991) Preferential utilization of specific immunoglobulin heavy chain diversity and joining<br />

segments in adult human peripheral blood B lymphocytes. J. Exp. Med. 173, 395– 407.<br />

13. Brisco, M. J., Condon, J., Sykes, P. J., Neoh, S. H., and Morley, A. A. (1991) Detection and<br />

quantitation of neoplastic cells in acute lymphoblastic leukaemia, by use of the polymerase<br />

chain reaction. Br. J. Haematol. 79, 211–217.<br />

14. Wiesner, R. J. (1992) Direct quantitation of picomolar concentrations of mRNAs by<br />

mathematical analysis of a reverse transcription/exponential polymerase chain reaction<br />

assay. Nucleic Acids Res. 51, 5863–5864.<br />

15. Taswell, C. (1981) Limiting dilution assays for the determination of immunocompetent<br />

cell frequencies. J. Immunol. 126, 1614–1619.<br />

16. Cremer, F. W., Dada, R., Kiel, K., Ehrbrecht, E., Goldschmidt, H., and Moos, M. (1998)<br />

Quantitative PCR monitoring of the effect of double high-dose therapy on the tumor load of<br />

patients with multiple myeloma. Ann. Hematol. 77 (Suppl. II), 182 (abstract 722).<br />

17. Vescio. R. A., Han, E. J., Schiller, G. J., Lee, J. C., Wu, C. H., Cao, J., et al. (1996)<br />

Quantitative comparison of multiple myeloma tumor contamination in bone marrow harvest<br />

and leukapheresis autografts. Bone Marrow Transplant. 18, 103–110.<br />

18. Cremer, F. W., Kiel, K., Wallmeier, M., Haas, R., Goldschmidt, H., and Moos, M. (1998)<br />

Leukapheresis products in multiple myeloma: lower tumor load after mobilization with<br />

cyclophosphamide plus granulocyte colony-stimulating factor (G-CSF) compared with<br />

G-CSF alone. Exp. Hematol. 26, 969–975.<br />

19. Kiel, K., Cremer, F. W., Ehrbrecht, E., Wallmeier, M., Hegenbart, U., Goldschmidt, H., aet<br />

al. (1998) First and second apheresis in patients with multiple myeloma: no differences in<br />

tumor load and hematopoietic stem cell yield. Bone Marrow Transplant. 21, 1109–1115.<br />

20. Kiel, K., Cremer, F. W., Rottenburger, C., Kallmeyer, C., Ehrbrecht, E., Atzberger, A., et al.<br />

(1998) Circulating tumor cells in patients with multiple myeloma in the course of high-dose<br />

therapy. Ann. Hematol. 77 (Suppl. II), 182 (abstract 724).<br />

21. Rottenburger, C., Kiel, K., Cremer, F. W., B′′sing, T., Atzberger, A., Moldenhauer, G.,<br />

et al. (1998) No differences in the tumor load in the CD19+ and CD20+ cell fractions of<br />

peripheral blood from patients with multiple myeloma post high-dose therapy and with<br />

peripheral blood stem cell support. Ann. Hematol. 77 (Suppl. II), 184 (abstract 731).


qPCR to Detect RNA Viruses 197<br />

32<br />

Ultrasensitive Quantitative PCR to Detect RNA Viruses<br />

Susan McDonagh<br />

1. Introduction<br />

The use of quantitative polymerase chain reaction (qPCR) to detect RNA viruses<br />

has become increasingly important as a prognostic marker and in patient management,<br />

for example, in human immunodeficiency virus (HIV) and hepatitis C virus (HCV)<br />

infection. Drug therapies can be monitored by regularly checking viral load, indicating<br />

whether the regime is sufficient, or whether alternatives should be sought. It is therefore<br />

crucial that the systems used are ultrasensitive and give accurate and reproducible<br />

results.<br />

There are several elements to consider in developing a reliable and accurate<br />

quantitative assay. First, rapid transport and storage of samples is important because<br />

of the unstable nature of RNA, as is the method of preparation. Samples should be<br />

received and processed within 6 h and the relevant fractions stored at –70°C until<br />

testing. Second, it is difficult to ascertain the efficiency of sample preparation methods;<br />

therefore, values obtained using qPCR may be several times lower than the actual copy<br />

number. Known standards should therefore be processed alongside samples to assess<br />

the loss within the system (1). Third, the efficiency of the reverse transcriptase step has<br />

been assessed to be from 5 (2) to 10% (3,4), and this must be taken into account when<br />

calculating viral load. Finally, there is a need to amplify all viral types equally, which<br />

can be a problem with the high mutation rate observed in RNA viruses. As a result,<br />

primers should be chosen from highly conserved noncoding regions. such as the 5′ noncoding<br />

region of HCV (5) and of enteroviruses (6).<br />

Several quantitative systems that use a variety of approaches have been developed<br />

and published. These include limiting dilution; the use of external standard curves;<br />

co-amplification of an internal reference template; and competitive PCR using an<br />

internal control. There are advantages and disadvantages with each of these systems,<br />

and the system of choice will depend on several factors. These methods have been used<br />

to detect as few as 40 copies of HCV (7), 64 and 4 copies of HIV per ml of plasma or<br />

serum, respectively (8,9), and less than 10 copies of enterovirus (10).<br />

1.1. Limiting Dilution<br />

The linear relationship between the amount of template and product is the basis of<br />

the limiting dilution method. However, this only occurs over a limited range; therefore,<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

197


198 McDonagh<br />

the dilution method is a semiquantitative system. Another consideration is that several<br />

reactions have to be run if Poisson distribution analysis is to be conducted using the<br />

limiting dilution method. Perhaps the most important consideration is that tube-to-tube<br />

variation is not controlled using this method (1).<br />

1.2. External Standard Curves<br />

This is based on the generation of an external standard curve where a dilution series<br />

of known amounts of template are carried out alongside each run.<br />

1.3. Coamplification of an Internal Reference Template<br />

The problem of tube-to-tube variation can be overcome by the coamplification of<br />

a single-copy cellular gene and the target sequence. This means that both templates<br />

should be affected by any variations of amplification efficiency. However, because<br />

different templates and reactions have different efficiencies, the relative amounts of<br />

product could vary. An added problem is that this system cannot be used to quantitate<br />

extracellular organisms and therefore is of no value in measuring viruses, including<br />

HIV and HCV, in plasma samples (1). As a result, this method will not be considered<br />

further.<br />

1.4. Competitive PCR Using an Internal Control<br />

The use of a competitive internal control that is similar to the target apart from<br />

length or the existence of a restriction site may alleviate the problems encountered with<br />

other methods of qPCR, such as co-amplification of an internal reference template (9).<br />

It will also act as a control for inhibition (10). Both the target and control DNA should<br />

amplify with the same efficiency because they compete equally under given conditions<br />

and the ratio of products will therefore remain constant throughout amplification (11).<br />

The point at which target and control DNA are equal can then be found by direct<br />

comparison using gel electrophoresis and densitometry or by enzyme immunoassay<br />

(EIA). This method has become the method of choice for quantitative assays, but it is<br />

expensive and time consuming because it requires several reactions per sample.<br />

To conduct a competitive PCR, it is necessary to create a template that is virtually<br />

the same as the product to be amplified but which can be identified during the detection<br />

stage. Primer-directed mutagenesis can be used to create a deletion, insert, or substitution<br />

at any point in the fragment (12,13). An outline of this procedure can be seen<br />

in Fig. 1. Two separate reactions are conducted, each using an external primer plus<br />

an internal primer containing the appropriate changes, for example, to create a novel<br />

restriction enzyme (RE) site. This can then be used as a competitive template, which<br />

can be differentiated from amplified wild-type product. The products of these reactions<br />

are purified, denatured, and renatured together then added to a PCR containing the<br />

outer primers. The resulting amplicon contains the mutated site and can be used directly<br />

or cloned to create as much template as required.<br />

The analysis of products is also important. Agarose gel electrophoresis, sometimes<br />

with densitometry, is easy to perform and is suitable for differentiating wild type from<br />

mutated products after digestion with the appropriate enzyme.


qPCR to Detect RNA Viruses 199<br />

Fig. 1. Construction of a competitive PCR template. Figure modified from ref. 12.<br />

2. Materials<br />

Unless stated, all chemicals are supplied by Sigma (Poole, UK), or Merck. All stock<br />

solutions up to and including the PCR step should be made using RNA free water.<br />

2.1. Reverse Transcription (see Note 1)<br />

1. dNTPs (1 mM; Pharmacia).<br />

2. Anti-Sense primer (100 ng/µL).<br />

3. RNasin (20–40 units/µL; Promega, Southampton, UK).<br />

4. Reverse transcriptase (RT) and 10× buffer, for example, AMV (Promega).<br />

5. Dimethyl sulphoxide (DMSO).<br />

6. Thermocycler (Hybaid, Teddington, UK).


200 McDonagh<br />

2.2. Basic PCR<br />

1. dNTPs (2 mM, 10× stock; Pharmacia; see Note 2).<br />

2. Taq or Pfu DNA polymerase and 10× buffer (Sigma, Promega; see Note 3).<br />

3. Forward and reverse primers (2 µM; 10× stock).<br />

4. Mineral oil (Sigma) or wax beads (Perkin–Elmer, Warrington, UK).<br />

5. Thermocycler (Hybaid, Teddington, UK).<br />

2.3. PCR to Create Mutated Competitor Fragments<br />

(e.g., Containing a Novel RE Site)<br />

1. Forward and reverse primers used in conventional PCR (2 µM).<br />

2. Internal forward and reverse primers containing bases to create novel RE site (2 µM).<br />

3. DNA purification system (Geneclean, <strong>Bio</strong> 101, US).<br />

4. PCR cloning system (TA cloning kit; Promega or Invitrogen; Leek, The Netherlands).<br />

5. RE enzyme (to digest mutated product) or method of differentiating between wild-type<br />

and mutated product, such as a probe.<br />

3. Methods<br />

3.1. Reverse Transcription<br />

1. Mix together the following reagents in an ice bath:<br />

7.5 µL of RNAse free water<br />

3 µL of DMSO<br />

100 ng of antisense virus-specific primer (in 1 µL)<br />

2 µL × 10 RT buffer<br />

3 µL of 1 mM dNTPs<br />

0.5 µL of RNAsin (20–40 units/µL)<br />

5 µL of RNA, prepared by extraction method<br />

1 µL of RT<br />

2. Incubate for 30–60 min at 42°C followed by 5 min at 95°C and 5 min at 4°C.<br />

3. Pulse spin tubes and add 5 µL to PCR reaction or store at –20°C until required (see<br />

Note 4).<br />

3.2. PCR<br />

3.2.1. Basic Reaction<br />

1. Make up a master mix containing 5 µL of buffer (×10); 5 µL of dNTP (2 mM, final<br />

concentration 0.2 mM); 5 µL of each primer (2 µM, final concentration 0.2 µM); and 1 U<br />

DNA polymerase (see Notes 2 and 3).<br />

2. Overlay with oil if not using a thermocycler with a heated lid, then add 5 µL of prepared<br />

cDNA through the oil.<br />

3. A typical cycling protocol consists of denaturation at 95°C for 30 s, annealing at 5°C below<br />

the primer melting point for 30 s, and extension at 68 to 74°C for 45 s (see Note 5).<br />

4. Carry out as few cycles as necessary because the reaction is only linear over a short range<br />

and this may distort the results.<br />

5. To further increase sensitivity, it is beneficial to use nested PCR. As a general rule, the<br />

primary round should consist of 20 to 25 cycles and the secondary round up to 20 cycles. It<br />

may also be necessary to add diluted primary product (1/10) to the secondary round.<br />

3.2.2. Limiting Dilution Method<br />

A limiting dilution of cDNA is performed in 10-fold steps, and this is then added<br />

to the PCR. The sensitivity of the assay must be known, and the amount of cDNA in


qPCR to Detect RNA Viruses 201<br />

the samples can then be extrapolated by finding the detection end point. A specific<br />

volume of cDNA may also be added to a number of replicate reactions to give a Poisson<br />

distribution (5,14,15).<br />

3.2.3. The Use of External Standard Curves<br />

A 10-fold dilution series covering from 1 to 10 5 molecules is used to produce a<br />

standard calibration curve. Amplified products can then be analyzed using the desired<br />

detection system, for example, by microwell capture so that optical density values can<br />

be compared with those of the standard curve (4).<br />

3.3. Competitive PCR Using an Internal Control<br />

3.3.1. Creation of a Mutated Template Containing a Novel RE Site<br />

1. Perform two PCR reactions, each containing an external primer plus the appropriate<br />

internal primer with mismatches to create the RE site. Approximately 20 to 30 cycles using<br />

a low annealing temperature will allow for mismatches (Fig. 1).<br />

2. Analyze the products by agarose gel electrophoresis, excise the bands, and recover the<br />

DNA using Geneclean following manufacturer’s protocols.<br />

3. Heat the combined products to 94°C for 1 min, then cool and allow to renature.<br />

4. Amplify the renatured products using the external primers and normal cycling conditions.<br />

5. Check for the mutated site by RE digest (3 µL of product, 1 µL of 10× buffer, 5 µL of<br />

dH2 2 O, 10 U/µL enzyme) at the appropriate temperature for 1 h. Analyze products by<br />

agarose gel electrophoresis. Alternatively, sequence the amplicon.<br />

6. If required, the products may be cloned into a PCR cloning vector for further manipulation<br />

or simply used as PCR template to create as much competitor as necessary.<br />

3.3.2. Quantitative PCR Using the Mutated Competitor<br />

1. Perform a series of 10-fold titrations of mutated fragment with known copy number<br />

(1–10,000 copies; see Note 6).<br />

2. Add each to a separate basic PCR along with the unknown quantity of DNA to be amplified<br />

and perform as few rounds as necessary (20–30 rounds).<br />

3. Run products on 1 to 2% agarose and analyze by densitometry (gel or photograph; see<br />

Fig. 2).<br />

4. The point at which wild-type and mutant intensity is equal identifies the amount of wild<br />

type DNA to within 1 log.<br />

4. Notes<br />

1. The RT and PCR stages may be performed in a single-tube format and also combined with<br />

hot start when wax beads are used to separate the stages (7).<br />

2. PCR DIG labeling mix must also be included if using EIA detection methods. Instead<br />

of adding combined dNTPs, add 200 µM of dATP, dCTP, dGTP, 190 µM of dTTP, and<br />

10 µM of Dig dUTP.<br />

3. Pfu has been found to generate higher product yields (4).<br />

4. cDNA produced in this reaction appears to be slightly unstable. Long-term storage leads<br />

to significant reductions in the amount of target DNA.<br />

5. Hot start PCR where a heating step at 95°C is required to activate the polymerase may<br />

increase amplification efficiency.<br />

6. Further titrations can be performed, although this is expensive and time consuming.<br />

7. The 118-bp fragment must also be taken into account when finding the point of equilibrium.<br />

Although it is possible to detect the end point of this reaction by eye (between lanes 6 and 7),<br />

it is more accurate to use densitometry.


202 McDonagh<br />

Fig. 2. Quantitative PCR products after RE digestion to reveal true products (522 bp) and<br />

mutated competitor products (404 and 118 bp). Lanes 1 through 8 contain 15 copies of template<br />

and decreasing amounts of mutated competitor (10,000, 2000, 1000, 200, 100, 20, 10, and<br />

1 copy, respectively). The reaction was based on the amplification of a 522-bp product and a<br />

mutated competitor that was identical except for the creation of a SmaI site 118 bp into the<br />

fragment (see Note 7).<br />

References<br />

1. Clementi, M., Menzo, S., Bagnarelli, P., Manzin, A., Valenza, A., and Varaldo, P. E. (1993)<br />

Quantitative PCR and RT-PCR in virology. PCR Methods Appl. 2, 191–196.<br />

2. Zhang, L. Q., Simmonds, P., Ludlam, C. A., and Leigh Brown, A. J. (1991) Detection,<br />

quantification and sequencing of HIV-1 from plasma of seropositive individuals and from<br />

factor VIII concentrates. AIDS 5, 675–681.<br />

3. Kaneko, S., Murakami, S., Unoura, M., and Kobayashi, K. (1992) Quantitation of hepatitis<br />

C virus by competitive polymerase chain reaction. J. Med. Virol. 37, 278–282.<br />

4. Whitby, K. and Garson, J. A. (1995) Optomisation and evaluation of a quantitative<br />

chemiluminescent polymerase chain reaction assay for hepatitis C virus RNA. J. Virol.<br />

Methods. 51, 75–88.<br />

5. Hawkins, A., Davidson, F., and Simmonds, P. (1997) Comparison of plasma virus loads<br />

among individuals infected with hepatitis C virus (HCV) genotypes 1, 2, and 3 by<br />

Quantiplex HCV RNA Assay versions 1 and 2, Roche Monitor Assay and an in-house<br />

limiting dilution method. J. Clin. Microbiol. 35, 187–192.<br />

6. Lauwers, S., Bissay, V., and Rombaut, B. (1997) Development of an enterovirus specific<br />

PCR method for the quantification of enterovirus genomes in blood of diabetes patients.<br />

Clin. Diagn. Virol. 9, 135–139.<br />

7. Whitby, K. and Garson, J. A. (1997) A single tube two compartment reverse transcription<br />

polymerase chain reaction system for ultrasensitive quantitative detection of hepatitis C<br />

virus RNA. J. Virol. Methods 66, 15–18.


qPCR to Detect RNA Viruses 203<br />

8. Kashanchi, F., Melpolder, J. C., Epstein, J. S., and Sadaie, M. R. (1997) Rapid and sensitive<br />

detection of cell-associated HIV-1 in latently infected cell lines and in patient cells using<br />

sodium-n-butyrate induction and RT-PCR. J. Med. Virol. 52, 179–189.<br />

9. Liuzzi, G., Chirianni, A., Clementi, M., Bagnarelli, P., Valenza, A., Cataldo, P. T., et al.<br />

(1996) Analysis of HIV-1 load in blood, semen and saliva: Evidence for different viral<br />

compartments in a cross-sectional and longitudinal study. AIDS 10, F51–F56.<br />

10. Mayerat, C., Burgisser, P., Lavanchy, D., Mantegani, A., and Frei, P. C. (1996) Comparison<br />

of a competitive combined reverse transcription PCR assay with a branched-DNA assay<br />

for hepatitis C virus. J. Clin. Microbiol. 34, 2702–2706.<br />

11. Gilliland, G., Perrin, S., Blanchard, K., and Bunn, H. K. (1990) Analysis of cytokine<br />

mRNA and DNA: Detection and quantitation by competitive polymerase chain reaction.<br />

Proc. Natl. Acad. Sci. USA 87, 2725–2729.<br />

12. Higuchi, R., Krummel, B., and Saiki, R. K. (1988) A general method of in-vitro preparation<br />

and specific mutagenesis of DNA fragments: study of protein and DNA interactions.<br />

Nucleic Acids Res. 16, 7351–7367.<br />

13. Vallette, F., Mege, E., Reiss, A., and Adesnik, M. (1989) Construction of mutant and<br />

chimeric genes using the polymerase chain reaction. Nucleic Acids Res. 17, 723–733.<br />

14. Simmonds, P., Zhang, L. Q., Watson, H. G., Rebus, S., Ferguson, E. D., Balfe, P., et al.<br />

(1990) Hepatitis C quantification and sequencing in blood products, haemophiliacs, and<br />

drug users. Lancet 336, 1469–1472.<br />

15. van Kerckhoven, I., Fransen, K., Peeters, M., de Beenhouwer, H., Piot, P., and van der<br />

Groen, G. (1994) Quantification of human immunodeficiency virus in plasma by RNA<br />

PCR, viral culture, and p24 antigen. J. Clin. Microbiol. 32, 1669–1673.


204 McDonagh


Site-Directed Mutation and PCR Mimics 205<br />

33<br />

Quantitative PCR for cAMP RI Alpha mRNA<br />

Use of Site-Directed Mutation and PCR Mimics<br />

<strong>John</strong> M. S. <strong>Bartlett</strong><br />

1. Introduction<br />

Precise and accurate determination of mRNA expression levels in tissues and model<br />

systems is a central methodology in a wide range of research applications. Expression<br />

of many genes is currently assessed by northern blotting, RNAse protection assays,<br />

Serial Analysis of Gene Expression (SAGE), and many other techniques; however, for<br />

single transcripts, especially where tissue is limited or abundance is low, quantitative<br />

polymerase chain reaction (PCR) is the method of choice. However, where quantitative<br />

PCR is to be used, the reproducibility, accuracy, and detection limits of the technique<br />

must be clearly defined.<br />

We address this issue in the context of breast cancer by establishing a quantitative<br />

PCR technique for the measurement of cAMP RI alpha-binding proteins, the regulatory<br />

subunitis of cAMP-dependent protein kinase, using PCR mimics. The technique was<br />

evaluated for interassay and intraassay variation and precision. This approach is<br />

applicable to any marker of interest where quantitation of RNA or DNA levels by<br />

PCR is attempted. For details of the construction of the mimics. see <strong>Bartlett</strong> et al. (1);<br />

however, a more robust and reproducible method of mimic construction is described<br />

within this volume (see Chapter 7).<br />

2. Materials<br />

1. RI alpha PCR primers (100 µM, see Note 1).<br />

RI alpha 430 sense: 5′-GCATAACATTCAAAGCACTGC-3′<br />

RI alpha antisense: 5′-CTTGCTGAATCACAGTCTCTCC-3′<br />

2. RI alpha control (430 base pairs) in pCRII vector: Dilute the control plasmid to 100, 10,<br />

1, and 0.1 pg of insert per µL (see Note 2).<br />

3. RI alpha MIMIC (430 base pairs) in pCRII vector (see Note 3): Dilute the mimic plasmid<br />

to 100, 10, 1, and 0.1 pg of insert per µL.<br />

4. Extracted RNA (1 µg in 5 µL per sample; see Subheadings 2.1.3.–2.1.5., ibid).<br />

5. Random hexamer N6 (100 µM, Life Technologies, Paisley, UK, see Note 4).<br />

6. dNTP mix: 2 mM each of dATP, dTTP, dCTP and dGTP in distilled water, aliquoted, and<br />

stored at –20°C (Life Technologies, Paisley, UK).<br />

7. MMLV Reverse transcriptase and buffer (200 units/µL; Life Technologies, Paisley, UK).<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

205


206 <strong>Bartlett</strong><br />

Table 1<br />

Example Assay Layout<br />

Sample<br />

Control 10 pg of RIα 100 pg of RIα Sample 1 Sample 2 Sample 3 Etc.<br />

0.1 pg of RIα<br />

mutant<br />

1.0 pg of RIα<br />

mutant<br />

10 pg of RIα<br />

mutant<br />

100 pg of RIα<br />

mutant<br />

8. Human placental ribonuclease inhibitor (40 U/µL, Pharmacia, UK).<br />

9. Radioactive dNTP mix: 1.25 mM each of dATP, dTTP, dGTP, and 0.5 mM dCTP in distilled<br />

water, aliquoted, and stored at –20°C (Life Technologies, Paisley, UK). Before use, add<br />

0.1 µCi 32 P dCTP (Amersham, UK)/5 µL dNTP mix to be used in PCR (see Note 5).<br />

10. Taq polymerase (5 units/µL) and buffer (Applied <strong>Bio</strong>systems, UK) 10× buffer: 500 mM<br />

potassium chloride, 100 mM Tris-HCl, 1% Triton-X and 25 mM magnesium chloride<br />

(see Note 6).<br />

11. Paraffin oil (see Note 7).<br />

12. Sodium chloride (100 mM) in sterile distilled water.<br />

13. EcoRV restriction enzyme 10 units/µL (Pharmacia UK).<br />

14. PAGE electrophoresis system (e.g., Protean II, <strong>Bio</strong>-Rad UK).<br />

15. Gel fixative: 5% glacial acetic acid, 40% methanol, 10% glycerol in water.<br />

16. Gel dryer.<br />

3. Methods<br />

3.1. Reverse Transcription<br />

1. Add 1 µL of random hexamer (100 ng) to 1 µg of RNA and make up to a total volume of<br />

9.5 µL with distilled water in a 0.2-mL thin-walled PCR tube.<br />

2. Heat to 65°C for 10 min and cool on ice.<br />

3. Prepare reverse transcription mix as follows: 4 µL of 5× reverse transcriptase buffer, 5 µL<br />

of 2 mM dNTP mix, 1 µL of reverse transcriptase (200 units), 0.5 µL of human placental<br />

ribonuclease inhibitor per reverse transcription reaction.<br />

4. Incubate at 42°C for 1 h, inactivate reverse transcriptase at 80°C for 5 min, and store<br />

cDNA at –20°C until required for quantitative PCR (see Note 8).<br />

3.2. Quantitative PCR<br />

1. Construction of standard reactions: Duplicate reactions were set up as follows: to separate<br />

aliquots of 10 pg of control RI alpha (in 10 µL), add 0.1, 1, 10, and 100 pg of mutant RI<br />

alpha (in 10 µL). Also, to separate aliquots of 100 pg of control RI alpha (in 10 µL), add<br />

0.1, 1, 10, and 100 pg of mutant RI alpha (in 10 µL). Prepare separate tubes only containing<br />

10 pg of mutant plasmid or 10 pg of control plasmid (restriction digest controls).<br />

2. For each sample (unknown concentration), take 1 µL of cDNA from the reverse transcription<br />

reaction plus 9 µL distilled water and to separate aliquots add 0.1, 1, 10, and 100 pg of<br />

mutant RI alpha (in 10 µL; see Table 1 for assay layout; also, see Note 9).


Site-Directed Mutation and PCR Mimics 207<br />

Fig. 1. Example of a crossover plot for a sample containing 10 pg of RI alpha cDNA. Vertical<br />

axis, counts per minute. Horizontal axis concentration of RI alpha mutant cDNA added. Solid<br />

curve = mutant RI alpha plasmid, dashed curve = sample. Reproduced with permission from<br />

<strong>Bartlett</strong> et al. (1).<br />

3. Prepare PCR master mix as follows: per tube, add To each tube 5 µL of 10× reaction<br />

buffer, 5 µL of 25 mM MgCl 2 , 0.1 µL of Taq polymerase, 5 µL of radioactive dNTP mix,<br />

and 14.9 µL of distilled water.<br />

4. Add 30 µL of PCR master mix to each tube set up in steps 1 and 2. If not using a heated<br />

lid PCR block, overlay with 100 µL of paraffin oil.<br />

5. Perform PCR amplification for 26 cycles (94°C for 40 s, 55°C for 60 s, 72°C for 70 s)<br />

followed by extension at 72°C for 5 min.<br />

3.3. Restriction Digestion<br />

1. Add 5.0 µL of 100 mM sodium chloride to each reaction followed by 1.0 µL of EcoRV<br />

restriction enzyme (10 units; see Note 10).<br />

2. Incubate at 37°C for 2 h to ensure complete digestion of mutant RIα product.<br />

3. Labeled PCR products were separated on a 6% polyacrylamide gel at 30 mA for 2 to 3 h<br />

using a Protean II vertical electrophoresis system (<strong>Bio</strong>-Rad UK).<br />

4. Gels were fixed for 30 to 60 min in fixative and dried using a flat bed gel dryer and heating<br />

to 80°C for 1 to 2 h.<br />

5. Gels were exposed to preflashed X-OMAT (Kodak UK) film for 1 to 8 h using radioactive<br />

ink to orient the gels.<br />

6. Bands corresponding to normal (430 base pair) and mutant component (215 base pair)<br />

component of each reaction were excised and 32 P incorporation determined by Cerenkov<br />

counting (see Note 11).


208 <strong>Bartlett</strong><br />

3.4. Calculation of Results<br />

1. The results of a typical assay are illustrated in Fig. 1. After restriction digestion of the<br />

co-amplified mutant and normal RI alpha, two bands are clearly visible, representing<br />

products of 430 and 215 base pairs. If no normal cDNA is added, all the product is digested<br />

to give only a 215-bp band. In the illustrated assay, mutant RI alpha cDNA is co-amplified<br />

in a range of concentrations with known (100 or 10 pg) or unknown (patient sample)<br />

concentrations of unmutated RI alpha cDNA. At high concentrations of mutant plasmid,<br />

the lower band is the most intense. As the concentration of mutant cDNA is decreased,<br />

the relative intensity of the lower 215 bp band decreases and that of the larger 430 bp<br />

band increases. Where concentrations are equivalent, each product is produced at the<br />

same intensity. Thereafter, the upper 430-bp band becomes more intense. The point of<br />

equivalence of concentration is therefore represented by a crossover between the lower<br />

and upper band intensities.<br />

2. For each sample, the counts per minute determined by Cerenkov (1) counting for the<br />

mutant and normal bands are plotted against the concentration of mutant plasmid added.<br />

The point at which the two curves cross represents the point at which the concentration of<br />

mutant and normal RI alpha are equivalent, thus allowing the concentration in unknown samples<br />

to be determined from this crossover point (see Fig. 2). In this example, the counts for<br />

the mutant and normal RI alpha bands from the PCR assay are plotted against the concentration<br />

of mutant RI alpha added. The curves cross over at 10 pg, indicating a concentration<br />

in the test sample of 10 pg RI alpha cDNA equivalent to 3.4 fmol RI alpha mRNA in<br />

the sample.<br />

3. The sensitivity of the assay was assessed using a range of cDNA concentrations from<br />

34 fmol to 0.0034 attomol (10 –14 to 10 –21 mol, 30 cycles of PCR were used for this lower<br />

limit). The sensitivity under these conditions was 0.002 attomol (2 × 10 –21 mol or approx<br />

1000 copies of mRNA; see Note 12).<br />

4. Intra and interassay variation for the quantitative PCR assay were determined as 8.0 and<br />

14.3%, respectively (see Note 13).<br />

5. Calculation of confidence intervals for results: As the co-efficients of variation are known<br />

for each stage of the reverse transcription (RT)-PCR assay, we were able to calculate<br />

the confidence intervals for sample concentrations determined by this assay technique.<br />

Where samples are measured within the same assay this is calculated as follows (see<br />

Note 13):<br />

C imax = C o (1 + E i )(1 + V i ) and C imin = C o (1 – E i )(1 – V i )<br />

Where C imax is the maximum estimated concentration and C imin is the minimum. C o is the<br />

observed concentration, E i is the intra-assay variation for reverse transcriptase, and V i is<br />

the intra-assay variation for PCR.<br />

If necessary interassay confidence limits can be defined from the above values using<br />

the following formula:<br />

C bmax = C imax (1 + E b )(1 + V b ) and C bmin = C imin (1 – E b )(1 – V b )<br />

Where C bmax is the maximum estimated concentration and C bmin is the minimum and E b<br />

is the inter-assay variation for reverse transcriptase and V b is the intra-assay variation<br />

for PCR.<br />

3.5. Discussion<br />

By assessing the variation at each step during the RT-PCR procedure, this method<br />

defines the variation as a result of the reverse transcription and PCR steps and show that


Site-Directed Mutation and PCR Mimics 209<br />

for samples assayed within a single assay the variation can be kept within acceptable<br />

limits. Furthermore, although it is possible, using the pCRII vector, to generate mRNA<br />

as an additional control, the low intra assay variation defined in this system allows this<br />

simpler procedure to be followed.<br />

Interassay and intra-assay variations were similar to those obtained by conventional<br />

radioimmunoassays (2,3), suggesting this technique could be robustly applied to<br />

clinical diagnostic problems, such as the determination of viral load for either RNA or<br />

DNA viruses (omitting the RT step for the latter).<br />

The sensitivity of this technique is such that low copy number genes could be<br />

assayed in 100 s or at most a few thousand cells and can be applied to patient tissue<br />

samples and small cell cultures. In addition, by allowing absolute concentrations to<br />

be determined, this assay will facilitate comparisons between laboratories previously<br />

hampered by semiquantitative approaches to PCR (4,5), and this precision is maintained<br />

even over most applications of fluorescence real-time PCR (6).<br />

4. Notes<br />

1. The cDNA sequence for human cAMP-dependent protein kinase subunit was retrieved<br />

from the Gembl database (accession no. M33336; 7). Using this sequence, primers were<br />

designed that amplified bases 159 to 589. This 430-bp fragment codes entirely for mRNA,<br />

which is subsequently translated into protein.<br />

2. The mimic is inserted in the 3900-bp pCRII vector by calculating the length of the vector<br />

containing the insert and dividing this by the length of the insert the molecular weight fraction<br />

made up by the insert is calculated and the Mwt corrected to reflect that of the PCR insert<br />

only. The molar equivalent for RI alpha mRNA added to the assay system was calculated<br />

as follows: 1 pg RI alpha plasmid is equivalent to 0.34 fmols RI alpha mRNA.<br />

3. The PCR mimic was constructed as described elsewhere. Using site-directed mutagenesis,<br />

a single bp change was introduced within the RI alpha PCR product to introduce a EcoRV<br />

restriction site. For researchers wishing to evaluate this quantitative PCR approach, this<br />

mimic is available on request from the author.<br />

4. Reverse transcription with a random hexamer will target all RNA species. Specific primers<br />

for the product of interest, out with the PCR product sequence, or polyA primers to target<br />

mRNA may be substituted as required.<br />

5. Because of the high activity of 32 P, alternatives such as 33 P or biotin-dCTP may be<br />

considered.<br />

6. Magnesium chloride concentrations may need to be altered for different PCR products. We<br />

therefore recommend using buffer with separate magnesium chloride.<br />

7. Using heated lids avoids the need for paraffin oil overlay, which reduces evaporation.<br />

8. To evaluate the reproducibility of the reverse transcription reaction we recommend<br />

including 0.5 µCi [ 35 S]dATP. This small amount of [ 35 S]dATP does not affect subsequent<br />

quantitation of the PCR. In our previous experiments, we demonstrated an intra-assay<br />

variation of 17% between samples in RT reactions and a 9% mean inter assay variation.<br />

9. The sensitivity of the assay can be varied by varying the standard range; decreasing the<br />

range of mutant and control plasmid from 0.1 to 100 pg to 0.01 pg to 1.0 pg increased the<br />

sensitivity of detection. For these experiments, PCR was performed over 30 cycles.<br />

10. The restriction enzyme EcoRV has an optimal digestion in buffered 50 mM KCl, 10 mM<br />

NaCl, and addition of 10% 100 mM NaCl to the PCR mix best approximates these<br />

conditions and avoids the need for purification of the PCR product at this step. The use of<br />

amplified control and mutant plasmids provides a high degree of control over the efficiency<br />

of the restriction digestion reaction.


210 <strong>Bartlett</strong><br />

11. Cerenkov counting is a simple principle where excised gel fragments are counted in the<br />

absence of scintillation fluid. Image analysis of the exposed autoradiograph can also be<br />

used at this stage.<br />

12. The assay has not been evaluated below this limit and theoretically could function at<br />

sensitivities of tens or hundreds of mRNA copies; at these low concentrations, stochastic<br />

errors may be introduced.<br />

13. Where all samples from a given individual or experiment are included in the same assay,<br />

interassay variation can be ignored.<br />

References<br />

1. <strong>Bartlett</strong>, J. M. S., Hulme, M. J., and Miller, W. R. (1996) Analysis of cAMP RI-alpha<br />

messenger-RNA expression in breast cancer—Evaluation of quantitative polymerase<br />

chain-reaction for routine use. Br. J. Cancer 73, 1538–1544.<br />

2. <strong>Bartlett</strong>, J. M., Wu, F. C. W., and Sharpe, R. M. (1987) Enhancement of Leydig cell<br />

testosterone secretion by isolated seminiferous tubules during co-perifusion in vitro:<br />

comparison with static co-culture systems. Intl. J. Androl. 10, 603–617.<br />

3. <strong>Bartlett</strong>, J. M. S., Weinbauer, G. F., and Nieschlag, E. (1989) Quantitative analysis of<br />

germ cell numbers and relation to intratesticular testosterone following vitamin a-induced<br />

synchronization of spermatogenesis in the rat. J. Endocrinol. 123, 403– 412.<br />

4. Touraine, P., Martini, J. F., Zafrani, B., Durand, J. C., Labaille, F., Malet, C., et al. (1998)<br />

Increased expression of prolactin receptor gene assessed by quantitative polymerase chain<br />

reaction in human breast tumors versus normal breast tissues. J. Clin. Endocrinol. Metab.<br />

83, 667–674.<br />

5. Sugimoto, T., Tsukamato, F., Fujita, M., and Takai, S. (1994) Ki-ras and c-myc oncogene<br />

expression measured by coamplification polymerase chain reaction. <strong>Bio</strong>chem. <strong>Bio</strong>phys.<br />

Res. Commun. 201, 574–580.<br />

6. Bieche, I., Olivi, M., Champeme, M. H., Vidaud, D., Lidereau, R., and Vidaud, M. (1998)<br />

Novel approach to quantitative polymerase chain reaction using real-time detection:<br />

application to the detection of gene amplification in breast cancer. Intl. J. Cancer 78,<br />

661–666.<br />

7. Sandberg, M., Skalhegg, B., and Jahnsen, T. (1990) The two mRNA forms for the type<br />

I alpha regulatory subunit of cAMP-dependent protein kinase from human testis are due<br />

to the use of different polyadenylation site signals. <strong>Bio</strong>chem. <strong>Bio</strong>phys. Res. Commun.<br />

167, 323–330.


Quantitation of Multiple RNA Species 211<br />

34<br />

Quantitation of Multiple RNA Species<br />

Ron Kerr<br />

1. Introduction<br />

Real-time quantitative polymerase chain reaction (PCR) methods (1) can be further<br />

optimized for various purposes by performing quantitative PCR for multiple RNA<br />

species in one sample (2). Advantages of this method not only include the elimination<br />

of differences in reaction mix volumes and conditions that may occur in separate<br />

samples, but importantly it also allows reference of RNA quantitation to an internal<br />

control (also see Notes 1 and 2). Internal controls are RNA species for which the<br />

quantity of RNA does not change across different cell types or conditions. They allow<br />

correction of RNA quantitation for different starting quantities of total RNA. Widely<br />

used examples of internal controls are 18S ribosomal RNA and glyceraldehyde-3-<br />

phosphate dehydrogenase (GAPDH) (3), and commercial standard kits for these are<br />

available. Different internal controls may be appropriate depending on the cell types<br />

or conditions being studied, and it is advisable to try several different controls when<br />

setting up a new assay. Quantitation of multiple RNA species is possible through the<br />

use of fluorescent probes, such as Taqman probes, as previously described.<br />

When multiple RNA species are to be quantified in a single sample, probes with<br />

fluorophores of different wavelengths are used. As with any assay that relies on<br />

the measurement of multiple fluorescent signals, it is important to ensure that the<br />

wavelengths of the emitted signals are as distant as possible to reduce overlap in<br />

signal detection (Fig. 1). The materials and methods for quantitation of mRNA for the<br />

fibrinolytic protein tissue plasminogen activator (t-PA) using 18S ribosomal RNA as<br />

an internal control are shown below as a worked example of the quantitation of two<br />

RNA species in a single sample.<br />

2. Materials<br />

All reagents may be obtained from Applied <strong>Bio</strong>systems, Warrington, Cheshire, UK,<br />

unless otherwise stated.<br />

1. 100 ng/µL RNA sample (see Chapters 9–11 for RNA isolation method).<br />

2. Reverse Transcriptase mastermix (9 µL per reaction): 2.85 µL of DEPC distilled H 2 O,<br />

1 µL of 10× Taqman buffer, 2.2 µL of 25 mM MgCl 2 , 2 µL of dNTPs, 0.2 µL of RNase<br />

Inhibitor, 0.25 µL of multiscribe RT, and 0.5 µL of random hexamers.<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

211


212 Kerr<br />

Fig. 1. Fluorescent probes 1 to 4 have different wavelengths as shown. For quantitation of<br />

multiple RNA species in a single sample, probes with the greatest difference in wavelengths<br />

should be selected to minimize overlap in signal detection. In this example probes 1 and 4<br />

would be selected.<br />

3. Primers/probes: Strict conditions must be adhered to for primer/probe design when using<br />

the ABI 7700 quantitative PCR system. These are shown in Table 1 and may be designed<br />

using Primer Express software. Primer and probe concentrations used are 10 pmol/µL<br />

and 5 pmol/µL, respectively.<br />

t-PA forward primer: GCAGGCTGACGTGGGAGTAC<br />

t-PA reverse primer: CCTCCTTTGATGCGAAACTGA<br />

t-PA probe: TGATGTGCCCTCCTGCTCCACCT<br />

These primers generate an amplicon of 91 base pairs, which spans intron 9 (1249 bp)<br />

of the t-PA gene. Because of the large intron size, only t-PA mRNA and not genomic<br />

DNA is amplified, removing the need for a DNase step. The t-PA Taqman probe has a<br />

FAM fluorescent label.<br />

4. 18S rRNA primer-probe mix (supplied premixed). The 18S rRNA probe has a VIC<br />

fluorescent label.<br />

5. PCR mastermix (Stratagene, La Jolla, California): 100 µL (500U) hot start Taq polymerase,<br />

1.7 mL of 10× PCR buffer, 1.44 mL of 50 mM MgCl 2 , 400 µL of 5 mM dNTPs, 6 µL<br />

of reference dye, and 6.354 mL of DEPC distilled H 2 O. This can be stored at –20°C<br />

until needed.<br />

6. Standard wall 0.6-mL capped conical tubes.<br />

7. 96-well reaction plate.<br />

8. ABI Prism 7700 Sequence detection system.<br />

3. Methods<br />

1. Reverse Transcriptase mastermix (9 µL) is added to 0.6-mL conical tubes.<br />

2. To this, 1 µL (approx 100 ng/µL) of total RNA is added.


Quantitation of Multiple RNA Species 213<br />

Table 1<br />

Conditions for Primer/Probe Design for Quantitative PCR<br />

Using the ABI 7700 Quantitative PCR System<br />

Primer<br />

T m (melting temperature) 58–60°C<br />

20–80% of nucleotides GC<br />

Length 9–40 bases<br />


214 Kerr<br />

after the PCR. This reduces the occurrence of contamination of subsequent assays<br />

by previous PCR products, which is an important consideration for any PCR system.<br />

A disadvantage of this system is that it may not be suitable for all target proteins.<br />

We have found that for some target RNAs trial of a large number of primer sets may be<br />

necessary before suitable primers are found. Also, the short amplicon length leads to an<br />

increased chance of homologous products amplified compared with PCR using primers<br />

for longer amplicons. We have found that the strict conditions may result in it not being<br />

possible to find suitable primers/probe for some RNA species. In our experience, these<br />

have been coagulation proteins, which are known to have a high degree of homology<br />

with other proteins.<br />

3. The above protocol will generate results as the cycle at which a set fluorescence threshold<br />

(Ct) is reached for each standard/unknown for both t-PA mRNA (FAM) and 18S rRNA<br />

(VIC). The standards are used to ensure that the difference between the Ct of the test RNA<br />

species and the Ct of the 18S rRNA internal control (∆Ct) remains constant across all<br />

mRNA dilutions in the range chosen for the standards (10 to 0.156× above). We can ensure<br />

that this is the case by plotting the ∆Ct for each of the standards against the log total RNA<br />

for each of the standards. If the gradient of the slope of this line is less than ± 0.1, this<br />

indicates that the ∆Ct is remaining constant and that the “∆∆Ct∀ method can be used<br />

to quantitate the mRNA for the unknowns as follows. The relative mRNA quantity of<br />

various unknowns compared with a nominal baseline unknown are calculated by using<br />

the formula 2 –∆∆ct , where ∆∆Ct is the difference in the ∆Ct of the test sample and the ∆Ct<br />

of the nominal baseline sample.<br />

The use of the ∆∆Ct method is advantageous when a large number of repeated samples<br />

have to be analyzed because it removes the need for running a set of standards each time,<br />

which is both time-consuming and increases costs. However, for many purposes, it seems<br />

sensible to generate a standard curve each time to ensure the validity of the results.<br />

References<br />

1. Heid, C. A., Stevens, J., Livak, K. J., and Williams, P. M. (1996) Real time quantitative<br />

PCR. Genome Res. 6, 986–994.<br />

2. Desjardin, L. E., Chen, Y., Perkins, M. D., Teixeira, L., Cave, M. D., and Eisenach, K. D.<br />

(1998) Comparison of the ABI 7700 system (TaqMan) and competitive PCR for quantification<br />

of IS6110 DNA in sputum during treatment of tuberculosis. J. Clin. Microbiol. 36,<br />

1964–1968.<br />

3. Goidin, D., Mamessier, A., Staquet, M. J., Schmitt, D., and Berthier-Vergnes, O. (2001)<br />

Ribosomal 18S RNA prevails over glyceraldehyde-3-phosphate dehydrogenase and betaactin<br />

genes as internal standard for quantitative comparison of mRNA levels in invasive and<br />

noninvasive human melanoma cell subpopulations. Anal. <strong>Bio</strong>chem. 295, 17–21.


Differential Display Techniques 217<br />

35<br />

Differential Display<br />

A Technical Overview<br />

<strong>John</strong> M. S. <strong>Bartlett</strong><br />

1. Introduction<br />

Since the completion of the human genome-sequencing project, scientists are now<br />

able to read the code of all human genes stored on the 46 chromosomes of the human<br />

genetic library. However, we are far from reaching an understanding of the functional<br />

relationships existing between more than a tiny fraction of these genes. The value of<br />

the human genome-sequencing project, beyond the simple collation of data, has been<br />

to teach us that it is possible to take a highly intensive analytical approach to the study<br />

of human systems in health and disease (1–4).<br />

The aim of our research has now shifted from the study of individual genes, isolated<br />

from the cellular environment in which they play their roles, to the investigation of the<br />

hugely complex interactions between gene and gene families. We are shifting from<br />

the observation of individual trees to an evaluation of the wood itself. To accomplish this,<br />

novel techniques have emerged that simultaneously allow the analysis of entire<br />

pathways or indeed entire cellular transcription patterns. The challenge this provides is<br />

both a molecular and mathematical one. Experiments now yield vast amounts of data<br />

that must be sifted through to formulate novel hypotheses. Conversely, we are now<br />

able to see how whole families of genes are regulated and interact rather than having to<br />

slowly piece together <strong>info</strong>rmation often from quite different experimental approaches.<br />

The continuing development of biomathematical modeling systems (5–9) is essential<br />

to this approach.<br />

Over the next decade, scientists face the challenge of transforming the knowledge<br />

gained of the human genome sequence into a practical and functional understanding<br />

of complex biological systems in health and disease. It is clear that analysis of gene<br />

expression represents a highly significant pointer to the altered function of transcripts<br />

identified by the human genome project whose function is largely unknown. The ability<br />

to select candidate genes from expression libraries from different tissues and disease<br />

states (and stages) for rapid investigation represents the single most important driver<br />

behind the current explosion in expression library analysis. It is therefore critical<br />

that we understand both the potentials and limitations of technologies available for<br />

expression analysis of entire transcriptomes.<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

217


218 <strong>Bartlett</strong><br />

As with any experimental approach, the researcher must have a clear goal and<br />

hypothesis in mind at the outset of the experimental procedure such that the techniques<br />

selected to achieve that goal are appropriate. The attraction, and pitfall, of transcriptome<br />

analysis is that these are extremely powerful tools for the identification of transcripts<br />

implicated in altered tissue functions. There are therefore a few basic principles that bear<br />

stating at this stage before the examination of the techniques available in this area.<br />

First and most importantly, a point commonly overlooked by those researching RNA<br />

expression is that proteins are effectors, and RNA is a blueprint. When examining RNA<br />

expression profiles, insight is gained into the regulation of expression and stability of<br />

the RNA species under examination, and these data must be extrapolated, usually by the<br />

assumption that expression of RNA relates to expression of protein. This extrapolation<br />

must be made with caution. However, the existence of the blueprint does not prove<br />

the material for which it codes has been produced and, conversely, failure to detect<br />

the template does not mean the protein is not present. It is therefore essential to link<br />

RNA analyses with analyses of protein expression and function; this requirement is<br />

frequently overlooked in transcriptome analyses.<br />

Second, in common with all analytical techniques, it is imperative that the composition<br />

of the tissue under examination be understood before correct interpretation of<br />

results can be performed. Although many transcriptome analyses are performed in<br />

monoclonal cell systems, frequently the analysis uses tissue, malignant or otherwise,<br />

as a starting point. Where mRNA is extracted from entire tissues, there exists a strong<br />

possibility of incorrectly assigning expression to the wrong cellular component. The<br />

presence of blood vessels, inflammatory cells, stromal components, and a mixture of<br />

diseased and normal tissue cells can complicate analyses. It is imperative, at some<br />

stage of the investigation, to determine the tissue source of the mRNA detected to avoid<br />

ascribing an incorrect origin to a protein or mRNA species.<br />

Each approach described within this section presents specific problems that must be<br />

sufficiently addressed to allow accurate and reliable data to be gathered if the effort<br />

expended is to be worthwhile. However, almost all techniques aimed at global analysis<br />

of RNA transcripts rely upon reverse transcription of mRNA, and most incorporate<br />

polymerase chain reaction (PCR). Reverse transcription (RT) comprises the transcription<br />

of the mRNA of choice into DNA (more accurately termed copy DNA or cDNA).<br />

This stage of the reaction is performed using RNA virus enzymes whose role in vivo<br />

is to transcribe viral RNA genome into a template for host transcription systems.<br />

The earliest RT enzymes used were derived from the Molony murine leukemia virus<br />

(Mo-MLV reverse transcriptase) and the avian myeloblastosis virus (AMV reverse<br />

transcriptase). More recently, genetically modified forms of these enzymes with<br />

enhanced activity have become the agents of choice. The enzymes are relatively heat<br />

labile (particularly in comparison with Taq polymerase) and reactions are performed at<br />

42°C with mRNA and nucleotides. The reaction must be primed because the enzymes<br />

are dependent on double-stranded nucleic acid template. The primer is usually allowed<br />

to anneal to the RNA after a short incubation (in the absence of enzyme) at 85°C to<br />

denature RNA tertiary structure. Primer selection is very important. Either a primer<br />

specific for the target sequence may be used or, more commonly, primers targeted at<br />

copying the entire mRNA population can be used. The use of poly (dT) primers will<br />

target mRNA poly A tails and transcribe from this point. In this case between 2 to 4 kb


Differential Display Techniques 219<br />

of cDNA can be routinely copied, which may place a constraint on PCR primer design.<br />

RNA can be transcribed with random hexanucleotide primers (more efficient than poly<br />

dT and produce cDNAs from the entire RNA pool); however, in this case, separation of<br />

mRNA from ribosomal and transfer RNAs is recommended as these will also be copied<br />

diluting the target sequences (10,11). The cDNA produced is then used for multiple<br />

PCRs using specific sequence primers. A further novel development has been the use<br />

of rTth DNA polymerase (from the bacterium Thermus thermophilus) allowing reverse<br />

transcription and PCR to be performed in a single reaction tube using this thermostable<br />

DNA polymerase with RT activity. Use of higher temperatures for the RT step with<br />

this system allows more efficient transcription, particularly from mRNAs with high<br />

guanine-cytosine content (12,13). The quality of mRNA extracted is an important<br />

confounder of many experimental approaches; even with conventional and proprietary<br />

mRNA extraction techniques, some carryover of polymerase or transcriptase inhibitors<br />

can be observed. In extreme cases, this may reduce the efficiency of the RT to an extent<br />

that causes errors in quantification, which may bias comparisons between experimental<br />

samples. Therefore, every effort should be made to ensure that mRNA preparations for<br />

quantification are as pure as possible.<br />

2. Systems for Transcriptome Analysis<br />

2.1. Differential Display<br />

The use of differential display technologies has rapidly expanded over the last<br />

decade as this technique has become established as a potent tool for the simultaneous<br />

analysis of multiple mRNA species (14). Differential display techniques have become<br />

as varied as the questions they are used to answer; however. in general they combine<br />

the following three separate techniques to address this single question.<br />

1. Production of cDNA from mRNA by RT.<br />

2. Design of arbitrary primers to allow parts of the cDNA (tags) to be amplified by PCR.<br />

3. Use of sequencing quality resolution by acrylamide electrophoresis.<br />

This approach builds up a fingerprint of the RNA species expressed in different<br />

tissue or cell samples. Comparison of these fingerprints identifies those genes that are<br />

upregulated or downregulated in different tissues. The technique is quite elegant in<br />

both its simplicity and power. However, as with many such techniques, the secret lies<br />

in the careful design and interpretation of the results.<br />

The model is best suited to the study of gene regulation in tissue culture where conditions<br />

can be varied under careful control to avoid artifacts. Even then extreme care<br />

must be taken to ensure that results are reproducible between experiments, the control<br />

of tissue culture conditions is in fact the most critical phase of the experiment in<br />

the production of accurate differential display results. Use of tissue samples makes the<br />

approach all the more complex. Although there is obvious value in investigations<br />

seeking to identify metastasis related gene expression, care must be taken to ensure<br />

that, for example, differences between premetastatic and postmetastatic tissues are<br />

related to metastasis and not a consequence of altered tumor–stroma interactions, or<br />

differences between primary tumors. In addition, extreme care must be taken to ensure<br />

samples to be compared are treated exactly the same to avoid artifacts being introduced<br />

during tissue storage, mRNA extraction, or subsequent amplification.


220 <strong>Bartlett</strong><br />

Even once novel regulatory changes are identified using differential display techniques,<br />

it remains common to double-check these changes using a second system, such<br />

as representational differential analyses (RDA), PCR. or Northerns, to confirm the<br />

result. This is an important confirmatory step before investing significant effort into the<br />

study of novel genes or gene transcripts identified by differential display (14). Even<br />

with these caveats, differential display has already proven itself as a highly valuable<br />

system for the study of gene expression changes during neoplastic progression.<br />

There are many similarities between differential display technologies and serial<br />

analysis of gene expression (SAGE). Indeed, SAGE might be seen as a logical progression<br />

from differential display. Both use degenerate primer sequences to produce a<br />

fingerprint of mRNA species for analysis. In differential display. these sequences may<br />

range from the highly selective (members of a particular gene family) to the more<br />

inclusive (polyT based) (14–21). Both methods use tagged primers, differential display<br />

for the purpose of detection, and SAGE for the purpose of capture and ligation. The<br />

fundamental difference between these approaches is that differential display continues<br />

to use a semiquantitative measure of expression and although more sensitive than<br />

microarrays in the detection of low copy number changes in expression, signal strength<br />

remains a determining factor of sensitivity. Again, identification of transcripts with<br />

altered expression relies on further analysis, either by sequencing of a representative<br />

clone or by blotting with a selective probe. A strength of this approach is illustrated<br />

by its ability to be targeted at specific genes or gene families. However, the number<br />

of transcripts able to be analyzed is limited by the electrophoretic separation required<br />

for the analysis of the results. Although this still has the capacity to recognize many<br />

hundreds of distinct transcripts, it is unlikely that the capacity of differential display<br />

can match that of either gene microarrays or SAGE. This section includes two different<br />

differential display protocols, one using targeted primers (AU Motifs) and the other<br />

being a more global protocol. Further techniques are available in an associated volume<br />

in this series (13).<br />

2.2. Microarrays<br />

DNA Chips or microarrays are becoming much more widely applied to the investigation<br />

of both model systems and whole tissues (14,22–25). However, these approaches<br />

are not, strictly speaking PCR related and therefore we have included an overview of<br />

arrays purely for completeness. DNA microarrays have the advantage of being data<br />

rich, that is to say it is possible to analyze many thousands of genes simultaneously.<br />

Using this approach it is possible to identify transcripts that are markedly upregulated<br />

or downregulated after experimentation or, indeed, as recently reported in the simple<br />

classification of cancers. This approach, in common with many conventional methods<br />

of gene expression analysis (northerns, RDA, etc.) relies on the measurement of<br />

signal intensity resulting from nucleic acid hybridization. Therefore, the efficiency<br />

of detection is a compound of the efficiency of labeling and hybridization of the<br />

individual clone. The resulting data give a semiquantitative estimate of changes in<br />

expression either up or down.<br />

The major weakness of the microarray approach is that it is firstly a semiquantitative<br />

approach and that it is therefore not optimal for the detection of low copy number<br />

gene transcripts. The major strength is the large number of genes (10 3 –10 4 ) that


Differential Display Techniques 221<br />

can be examined simultaneously in a single experiment and the rapidity with which<br />

<strong>info</strong>rmation can be gathered from multiple samples.<br />

2.3. SAGE<br />

In SAGE, short expressed sequence tags are produced from each mRNA expressed.<br />

Thus far, the technique is very similar to that of expressed sequence tag (EST) library<br />

screening. The advantage of SAGE is that the tags used are 9- to 10-bp long, and<br />

through a number of steps (see chapter 40 by Oien, these tags are concatenated into<br />

clones containing multiple (up to 40+) sequence tags before sequencing. Therefore, the<br />

expression profile is generated by sequencing many such clones and simple counting<br />

of the number of times each tag is represented (26).<br />

Therefore, the chief difference between DNA microarrays, differential display, and<br />

SAGE is that SAGE produces a digital count of the number of clones representing<br />

each sequence expressed. This is, however, only one of the differences between these<br />

approaches. Microarrays can identify large numbers (10,000 or more) of expressed<br />

genes, and determine, semiquantitatively, alterations in their expression. However,<br />

genes whose expression is modulated only marginally are unlikely to be identified by<br />

this approach. In SAGE, each expressed gene may be sequenced and thus single copy<br />

changes in expression may be detected if sufficient clones are sequenced. In SAGE,<br />

the sensitivity is determined not by the detection system but by the amount of effort<br />

given to identification of sequence tags. A further advantage of SAGE analysis over<br />

microarrays is that it is not necessary to have to hand a clone representing the gene(s) of<br />

interest; because the detection of genes is performed by sequencing, no hybridization<br />

matrix is required. SAGE analysis is therefore limited only by the amount of sequencing<br />

that can be performed in a cost-effective manner. Conventional sequencing of ESTs<br />

relies on the analysis of several hundred bps to identify each gene. In SAGE the<br />

sequence tag used to identify the expressed sequence is 10 bp. Theoretically, the<br />

discriminatory power of a 10-bp region of cDNA is 1 in 1,048,576 (4 10 ). By linking<br />

together the sequence tags from many different genes into a concatemeric clone, a<br />

single sequencing run of 800 bp can be used to identify 50 or more different sequences.<br />

Given the high throughput available in today’s 96-lane sequencing platforms, approx<br />

5000 gene transcripts may be analyzed from a SAGE library in a single sequencing<br />

run. In addition, SAGE has the advantage that digital data is more suited to statistical<br />

analysis. Finally, and perhaps of greatest significance as increasingly experiments<br />

become more time consuming and costly to perform the quantitative nature of SAGE<br />

library analysis potentially allows direct comparisons between laboratories and between<br />

libraries, subject only to the constraints of the initial experimental variables. Initial data<br />

comparisons support the premise that comparisons between laboratories performing<br />

SAGE can provide valuable and valid <strong>info</strong>rmation. The caveat that must be applied<br />

is that although such comparisons are at present validated by internal controls and<br />

performed in expert laboratories with rigorous controls, there are also specific pitfalls in<br />

SAGE analysis relating to identification of clones and PCR biases in construction of tags<br />

(26). Although some groups are already extrapolating from this data to compare data<br />

from different experimental approaches (27), in the light of some concerns raised and<br />

the possible impact on gene expression of even small changes in pH, nutrient, etc., this<br />

approach is at present premature, not just for SAGE but for all expression analyses.


222 <strong>Bartlett</strong><br />

Notwithstanding these caveats, SAGE analysis is a highly powerful experimental<br />

tool on which is increasingly applied to transcriptome analysis. The power of SAGE<br />

analysis is the product of an extremely complex experimental protocol, one that relies<br />

on the construction of a representative cDNA library for each RNA sample to be<br />

analyzed. In addition, the number of clones that must be sequenced for even a simple<br />

comparison between two libraries, despite the use of short sequence tags, is high<br />

(between 800–4000 has been suggested). Furthermore, although there is a high<br />

probability that a short sequence tag of 10 bp will identify a unique sequence, it is<br />

undeniably more complex to identify gene transcripts from such tags. Proponents<br />

of SAGE point out, with some justification, that the effort may well repay the cost.<br />

Nonetheless, the prospect of performing even a limited analysis of clinical material<br />

(50–100 tumors) using this method is a daunting one and many will be tempted to take<br />

a simpler, albeit less quantitative approach, such as microarray or differential display<br />

analysis, for their first steps in transcriptome analysis. More importantly, SAGE is<br />

an undirected technique that produces a global transcriptome map. Both microarrays<br />

and differential display techniques can, however, be readily tailored to the particular<br />

experimental question and hypothesis under investigation. As discussed at the outset of<br />

this chapter, careful consideration will be required to select the method most appropriate<br />

to the research question, expertise, and resources available to the investigator.<br />

References<br />

1. Zhang, L., Zhou, W., Velculescu, V. E., Kern, S. E., Hruban, R. H., Hamilton, S. R., et al.<br />

(1997) Gene expression profiles in normal and cancer cells. Science 276, 1268–1272.<br />

2. Lennon, G. G. (2000) High-throughput gene expression analysis for drug discovery. Drug<br />

Discovery Today 5, 59–66.<br />

3. Loging, W., Lal, A., Siu, I.-M., Loney, T., Wikstrand, C., Marra, M., et al. (2000) Identifying<br />

potential tumour markers and antigens by database mining and rapid expression screening.<br />

Genome Res. 10, 1393–1402.<br />

4. Szallasi, Z. (1998) Gene expression patterns and cancer. Nat. <strong>Bio</strong>technol. 16, 1292–1293.<br />

5. Larsson, M., Stahl, S., Uhlen, M., and Wennborg, A. (2000) Expression profile viewer<br />

(ExProView): A software tool for transcriptome analysis. Genomics 63, 341–353.<br />

6. Lash, A. E., Tolstoshev, C. M., Wagner, L., Schuler, G. D., Strausberg, R. L., Riggins,<br />

G. J., et al. (2000) SAGEmap: A public gene expression resource. Genome Res. 10,<br />

1051–1060.<br />

7. http://www.ncbi.nlm.nih.gov/geo/.<br />

8. www.ncbi.nlm.nih.gov/sage.<br />

9. http://www.ebi.ac.uk/arrayexpress/.<br />

10. Ji, H. J., Zhang, Q. Q., and Leung, B. S. (1990) Survey of oncogene and growth factor/<br />

receptor gene expression in cancer cells by intron-differential RNA/PCR. <strong>Bio</strong>chem. <strong>Bio</strong>phys.<br />

Res. Commun. 170, 569–575<br />

11. Jindal, S. K., Ishii, E., Letarte, M., Vera, S., Teerds, K. J., and Dorrington, J. H. (1995)<br />

Regulation of transforming growth factor alpha gene expression in an ovarian surface<br />

epithelial cell line derived from a human carcinoma. <strong>Bio</strong>l. Reprod. 52, 1027–1037.<br />

12. White, B. A., ed. (1984) PCR Protocols: Current Methods & Applications.Volume 15,<br />

Methods in Molecular <strong>Bio</strong>logy. Humana Press, Totowa, NJ.<br />

13. Liang, P. and Pardee, A. B., eds. (1998) Differential Display Methods & Protocols.Volume<br />

85, Methods in Molecular <strong>Bio</strong>logy. Humana Press, Totowa, NJ.<br />

14. Jurecic, R. and Belmont, J. W. (2000) Long-distance DD-PCR and cDNA microarrays.<br />

Curr. Opin. Microbiol. 3, 316–321.


Differential Display Techniques 223<br />

15. Marguiles, E. and Innis, J. (2000)eSAGE: Managing and analysing data generated with<br />

serial analysis of gene expression (SAGE). <strong>Bio</strong><strong>info</strong>rmatics 16, 650–651.<br />

16. Liang, P. and Pardee, A. B. (1998) Differential display: A general protocol. Mol. <strong>Bio</strong>technol.<br />

10, 261–267.<br />

17. Jorgensen, M., Bevort, M., Kledal, T. S., Hansen, B. V., Dalgaard, M., and Leffers, H.<br />

(1999) Differential display competitive polymerase chain reaction: An optimal tool for<br />

assaying gene expression. Electrophoresis 20, 230–240.<br />

18. Matz, M. V. and Lukyanov, S. A. (1998) Different strategies of differential display: Areas<br />

of application. Nucleic Acids Res. 26, 5537–5543.<br />

19. Lapointe, J. and Labrie, C. (1999) Identification and cloning of a novel androgen-responsive<br />

gene, uridine diphosphoglucose dehydrogenase, in human breast cancer cells. Endocrinology<br />

140, 4486–4493.<br />

20. Cirelli, C. and Tononi, G. (1999) Differences in brain gene expression between sleep<br />

and waking as revealed by mRNA differential display and cDNA microarray technology.<br />

J. Sleep Res. 8, 44–52. Supplement.<br />

21. Aittokallio, T., Ojala, P., Nevalainen, T. J., and Nevalainen, O (2000) Analysis of similarity<br />

of electrophoretic patterns in mRNA differential display. Electrophoresis 21, 2947–2956.<br />

22. Lockhart, D., Dong, H., Byrne, M., Follettie, M., Gallo, M., Chee, M., et al. (1996)<br />

Expression monitoring by hybridization to high-density oligonucleotide arrays. Nat.<br />

<strong>Bio</strong>technol. 14, 1675–1680.<br />

23. Schena, M., Shalon, D., Davis, R., and Brown, P. (1995) Quantitative monitoring of gene<br />

expression patterns with a complementary cDNA microarray. Science 270, 467– 470.<br />

24. Lomax, M. I., Huang, L., Cho, Y., Gong, T.-W., and Altschuler, R. A. (2000) Differential<br />

display and gene arrays to examine auditory plasticity. Hearing Res. 147, 293–302.<br />

25. Sun, Y. (2000) Identification and characterization of genes responsive to apoptosis:<br />

Application of DNA chip technology and mRNA differential display. Histol. Histopathol.<br />

15, 1271–1284.<br />

26. <strong>Bartlett</strong>, J. M. S. (2001) Technology evaluation: SAGE, Genzyme molecular oncology.<br />

Curr. Opin. Mol. Ther. 3, 85–96.<br />

27. Kal, A. J., Van Zonneveld, A. J., Benes, V., Van den Berg, M., Koerkamp, M. G., Albermann,<br />

K., et al. (1999) Dynamics of gene expression revealed by comparison of serial analysis<br />

of gene expression transcript profiles from yeast grown on two different carbon sources.<br />

Mol. <strong>Bio</strong>l. Cell 10, 1859–1872.


224 <strong>Bartlett</strong>


AU-Differential Display 225<br />

36<br />

AU-Differential Display, Reproducibility<br />

of a Differential mRNA Display Targeted to AU Motifs<br />

Orlando Dominguez, Lidia Sabater, Yaqoub Ashhab,<br />

Eva Belloso, and Ricardo Pujol-Borrell<br />

1. Introduction<br />

AU-rich elements (AREs) are found in 3′ untranslated regions (3′ UTR) of many<br />

highly unstable mRNAs for mammalian early-response genes. The minimal AU<br />

sequence core within the ARE is the heptamer WAUUUAW, although from a functional<br />

point of view, several pentanucleotides clustered in close proximity are the key<br />

sequence motif that mediates mRNA degradation (1).<br />

Genes containing AREs are of potential biological and pharmacological interest<br />

because they often code for inflammatory mediators, cytokines, proto-oncoproteins,<br />

and transcription factors (2–5).<br />

A targeted differential display named AU-motif directed display (AU-DD) is<br />

described here. It allows the isolation of cDNA fragments from ARE-containing<br />

mRNAs with minimal sequence <strong>info</strong>rmation. AU-DD combines a high specificity<br />

gained at the polymerase chain reaction (PCR) level with the advantages of direct<br />

comparison between samples at different stages of activation to detect differentially<br />

expressed genes (6,7).<br />

Previous examples of targeted differential display have used longer and wellconserved<br />

motifs from multigene families. These include zinc finger motifs (8,9),<br />

plant MADS boxes (10), and motifs specific for heat shock proteins (11). They employ<br />

longer conserved sequences (from 6 to 8 codons) to design primers that can predictably<br />

work well on PCR.<br />

The PCR step of AU-DD uses primers which contain two distinctive domains, a<br />

double AU motif at the 3′ end (AU2 in short) and a guanosine rich 5′ (Table 1). This<br />

primer core sequence is previously incorporated as a 5′ tag in the retrotranscription<br />

primer (see below). The 5′ domain configures a sticky anchor that increases Tm and<br />

stabilizes primer annealing whereas the 3′ confers motif specificity. Both 5′ and 3′<br />

domains contribute in a cooperative way to primer annealing. When the natural primerbinding<br />

site was identified on cloned products, it was always found to be a true AU-rich<br />

site. The use of the AU2 domain-based primers does not restrict; however, the spectrum<br />

of amplified mRNAs to clustered AU-pentamer containing genes. In fact, cDNAs<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

225


226 Dominguez et al.<br />

Table 1<br />

Primer Sequences for Both AU-DD and Control Genes<br />

Protocol Primer Designation Sequence<br />

(A) Retrotranscription (RT) and PCR Primers Used in AU-DD<br />

1 G7AU2dT (RT) GGGGGGGTATTTATTTA(ACGT)TTTTTTTTT<br />

TTTTTT(ACG)<br />

2 G7AU2 GGGGGGGTATTTATTTA<br />

G7AU2A<br />

GGGGGGGTATTTATTTAA<br />

G7AU2C<br />

GGGGGGGTATTTATTTAC<br />

G7AU2G<br />

GGGGGGGTATTTATTTAG<br />

G7AU2T<br />

GGGGGGGTATTTATTTAT<br />

3 GTGAU2dT (RT) GGTGGGTGGTATTTATTTA(ACGT)TTTTTTT<br />

TTTTTTTT(ACG)<br />

4 GTGAU2 GGTGGGTGGTATTTATTTA<br />

GTGAU2A<br />

GGTGGGTGGTATTTATTTAA<br />

GTGAU2C<br />

GGTGGGTGGTATTTATTTAC<br />

GTGAU2G<br />

GGTGGGTGGTATTTATTTAG<br />

GTGAU2T<br />

GGTGGGTGGTATTTATTTAT<br />

(B) Primers Used in Control PCR Experiments<br />

IFNα (J00207) IFNα513s GGCCTTGACCTTTGCTTTA<br />

IFNα915as<br />

CTTCATCAGGGGAGTCTCTGT<br />

GAPDH (M33197) GAPDH19s TCTTCTTTTGCGTCGCCAG<br />

GAPDH390as<br />

AGCCCCAGCCTTCTCCA<br />

(A) 1 and 3 are alternative retrotranscription (RT) primers; 2 and 4 list two series of PCR primers compatible<br />

with RT primers 1 and 3, respectively. (B) Specific primers for IFNA and GAPD, respectively.<br />

carrying single AU motifs were often amplified, and only 15% of cloned cDNAs<br />

contained a complete double AU element (12).<br />

On its part, the reverse transcriptase primer incorporates a significant feature by<br />

including the PCR primer as a 5′ tag. This allows the use of a single primer in the<br />

PCR, providing the possibility of a fine titration of the annealing temperature to<br />

avoid primer artifacts generated when two oligonucleotides are used in low stringency<br />

conditions.<br />

The cDNA is synthesized from total RNA trying to ensure initiation from true<br />

poly-A tails of mRNA, which is not always easy. With our protocol, only 8% of cloned<br />

products lacked the poly-A tail (12).<br />

For two mRNA species to be effectively sampled by DD, the following two conditions<br />

are critical: (1) comparable amounts of template must be used and (2) cDNA<br />

concentration must be above a threshold to obtain reproducible results. During all<br />

the initial processes, RNA concentration and integrity and the absence of contaminating<br />

DNA has to be monitored carefully. Even trace amounts of genomic DNA can<br />

significantly alter the final profile, yielding fake products and turning the technique<br />

unreliable. RNA should be checked for quantity and integrity both after purification<br />

and after DNase treatment. The use of good-quality DNase does not guarantee the<br />

removal of all DNA, and this process requires also monitoring. The assessment of


AU-Differential Display 227<br />

DNase treatment efficiency and titration and normalization of cDNA samples were<br />

performed by PCR as described (12).<br />

In general, it is always easy to get amplification products. However, this does<br />

not guarantee that those products are specific and reproducible. When the cDNA<br />

concentration is below a minimal threshold, the final products belong to nonspecific<br />

templates that were selected and amplified. This is why both minimal amounts of RNA<br />

to be retrotranscribed and cDNA normalization are fundamental. The minimal amount<br />

ensures that specific targets can be easily found and selected during the first cycles<br />

of PCR, and the normalization ensures that DD profiles between samples activated<br />

and resting can be compared.<br />

The major procedures described in the protocol refer to RNA preparation and DNase<br />

treatment, cDNA synthesis, AU-DD, and confirmation of differential expression.<br />

2. Materials<br />

In general, any reagent of analytical grade was considered of sufficient purity for<br />

general procedures, such as electrophoresis or initial stages of RNA purification.<br />

Otherwise, reagents added to enzymatic reactions or intended to dissolve RNA were of<br />

molecular biology or ultrapure grades.<br />

2.1. RNA Preparation<br />

1. Lysis solution: GCS (Guanidinium thiocyanate (GuTh), Citric acid, Sarkosyl) is a<br />

modification of Sacchi and Chomczynski solution D (14), that is, 4 M GuTh, 50 mM<br />

sodium citrate (citric acid adjusted to pH 3.8 with NaOH), and 0.5% Sarkosyl. The solution<br />

can be stored for up to no more than 1 mo at 4°C.<br />

2. Double-distilled phenol (Sigma). Phenol was saturated and equilibrated in ultrapure water<br />

and treated with 8-hydroxyquinoline (Sigma) (13) yielding unbuffered acidic phenol<br />

(pH 4 to 5). Aliquots containing a water overlay were dispensed in sterile polypropylene<br />

tubes and stored frozen at –20°C. After thawing, they were kept at 4°C and discarded<br />

after 2 wk.<br />

3. Complete Lysis solution (GCSMP): Mix GCS containing beta-2 mercaptoethanol up to 2%<br />

with 1 volume of acidic phenol to give GCSMP immediately before use.<br />

4. Phase lock heavy gel (PLG, Eppendorf) stored at room temperature.<br />

5. Glycogen as coprecipitant (Roche Molecular <strong>Bio</strong>chemicals) stored at –20°C.<br />

6. Chloroformisoamyl alcohol: Chloroform should be mixed with isoamyl alcohol in a<br />

proportion of 982 v/v (CI), respectively, in a Pyrex bottle with minimal air volume<br />

and kept at 4°C.<br />

7. DNAse I RQ1 from Promega at 0.2 U/µL in a solution containing 10 mM Bis-Tris-HCl,<br />

pH 6.5, 1 mM EDTA, 5 mM MgCl 2 , 5 mM DTT, and 1 U/µL RNAsin.<br />

8. RNasin Ribonuclease Inhibitor (40 U/µL) from Promega stored at –20°C.<br />

9. Ethanol 96% and 75%.<br />

10. 3 M Potassium Acetate pH 5.0 (KAc).<br />

11. Isopropanol.<br />

12. Guanidinium thiocyanate (4 M) in water.<br />

2.2. cDNA Synthesis and Further Treatment<br />

1. Reverse transcriptase: SuperScript II RNase H- (Gibco-BRL). Store at –20°C.<br />

2. dNTPs (Amersham <strong>Bio</strong>sciences) at 10 mM each.<br />

3. RNasin Ribonuclease Inhibitor 40 U/µL from Promega (Madison, WI) stored at –20°C.


228 Dominguez et al.<br />

4. Oligonucleotides used as reverse transcription primers (HPLC purified; Genset, Paris,<br />

France) are listed in Table 1. There are two variants used in independent but complementary<br />

experiments, G7AU2dT and GTGAU2dT.<br />

5. RNase H (Gibco BRL) stored at –20°C.<br />

6. Qiaquick PCR purification kit columns (Qiagen).<br />

7. EE (10 mM EPPS), 0.1 mM EDTA, pH 8.2 adjusted with 0.1 N NaOH was used throughout<br />

as a general elution/dilution solution.<br />

8. Reagents for agarose gel electrophoresis.<br />

9. GAPDH primers (see Table 1).<br />

2.3. AU-DD<br />

1. α- 32 P-dATP at 3000 Ci/mmol/DuPont NEN.<br />

2. Platinum Taq DNA polymerase (Gibco-BRL).<br />

3. PCR buffer 1×: 10 mM Tris-HCl, pH 8.8, 1.5 mM MgCl 2 , 50 mM KCl, 0.1% Triton<br />

X-100.<br />

4. dNTPs at 2 mM.<br />

5. AU-DD primers are listed in Table 1.<br />

6. For confirmation of differential expression by standard PCR, specific primers were<br />

derived from sequences of cloned AU-DD products with the program Oligo v5.0 (National<br />

<strong>Bio</strong>sciences Inc.) to have a Tm of 63 to 65°.<br />

2.4. Gel Electrophoresis<br />

1. S2 GibcoBRL electrophoresis apparatus for 0.8-mm thick gels.<br />

2. 40% (382) Polyacrylamide-bispolyacrylamide (Serva, Germany).<br />

3. TEMED (N,N,N′,N′-tetramethylethylethlenediamine).<br />

4. 10% Ammonium persulfate.<br />

5. Whatman 3 MM paper (gel blotting paper).<br />

6. TBE buffer (89 mM Tris Base, 89 mM boric acid, 1 mM EDTA).<br />

7. 20% methanol-10% acetic acid.<br />

3. Method<br />

A schematic representation of AU-DD method is shown in Fig. 1.<br />

3.1. RNA preparation (see Note 1)<br />

1a. Solid tissue is pulverized in liquid nitrogen with mortar and pestle (see Note 2).<br />

1b. Cultured cells are pelleted at 800g for 5 min at 4°C, and put on ice as dry pellets.<br />

1c. For culture flasks with adherent cells, decant, rinse with ice-cold phosphate-buffered saline<br />

and decant by aspiration.<br />

2. Add GCSMP in a proportion of 2 mL per 50 mg of tissue or 10 7 cells.<br />

3. Homogenize lysates at 25,000 rpm for 30 s with a mechanical homogenizer (Ultraturrax<br />

T25, Ika) to ensure both lysis and a complete DNA shearing. Stand for 2 min at room<br />

temperature (see Note 3).<br />

4. Add 0.4 volumes of CI (chloroformisoamyl alcohol) with respect to the complete lysis<br />

solution—including phenol volume—and shake energetically for 10 s.<br />

5. Pour on a prepacked PLG tube. Empty prepacked PLG tubes are 14-mL PPN round-bottom<br />

centrifuge tubes containing 1.5 mL of PLG-heavy and centrifuged 2 min at 1500g; it<br />

should then be left for 10 min on ice.<br />

6. Spin 10 min at 2000g on bench-top centrifuge with swinging bucket rotor. Save upper<br />

aqueous phase to high speed tubes and discard PLG interphase and bottom phase.


AU-Differential Display 229<br />

Fig. 1. Schematic representation of AU-DD. (1) Retrotranscription. Dotted line represents<br />

mRNA template. The RNA molecule at the top contains an AU-rich sequence (AU) and a poly-A<br />

tail. G-AU2TTTTTV represents a generic RT primer, with a 5′ anchor of rich on Guanosines,<br />

an AU2 sequence and an oligo(dT). (2) First round of PCR with a single primer GNAU2. (3)<br />

Final products of AU-DD amplification.<br />

7. Precipitate by mixing with 1 volume of isopropanol. Incubate 1 h at –20°C and centrifuge<br />

30 min at 10,000g. Wash with 75% ethanol.<br />

8. Dissolve pellets in 4 M guanidinium thiocyanate. Use at least 0.2 volumes of the lysis<br />

solution used initially. Heat if necessary (pulses of 5 min at 60°C) until there are<br />

not pellet particles left in solution. Be careful because the pellet particles become<br />

transparent and it is difficult to see them. Coprecipitant may be incorporated here adding<br />

1 µL of glycogen (20 mg/mL) and mix.<br />

9. Repeat the precipitation step with isopropanol (9).<br />

10. Wash with 75% ethanol. Dissolve in ultrapure water. The volume of water depends on the<br />

RNA concentration desired. Typically, 50 µL is appropriate to run out all the following<br />

experiments.<br />

11. Total RNA concentration is measured by spectrophotometry and gel electrophoresis. Not<br />

less than 3 µg (see Note 4) or up to 50% of RNA prep is DNase I treated (next step) in a<br />

separate tube, storing the other half as backup or for other purposes at –80°C after mixing<br />

with 3 volumes of 95% ethanol (see Note 5).<br />

12. RNA integrity and concentration is assessed by titration using Escherichia coli rRNAs<br />

(Sigma) as standard in 1% TBE agarose gel electrophoresis and ethidium bromide staining<br />

(13). Different concentrations of commercial E. coli rRNAs (800, 400, 200 ng) are<br />

compared in intensity with different dilutions of the RNA sample.<br />

13. Typically, 5 to 10 µg of total RNA (see Note 6) is incubated at 0.25 µg/µL with DNAse I<br />

for 30 min at 37°C. One microliter is then taken to check the absence of DNA (see<br />

Note 7).<br />

14. RNA is precipitated by adding 0.1 volumes of 3 M KAc, pH 5.0; 1 µL of glycogen; and<br />

3 volumes of 95% ethanol (see Note 8). After standing for 5 min at room temperature,<br />

the pellet is recovered by 10 min of centrifugation in a microcentrifuge at 10,000g and<br />

washed in 75% ethanol (see Note 9).


230 Dominguez et al.<br />

Fig. 2. cDNA normalization for their concentration in GAPDH gene. Lanes 1, 3, 5, and<br />

7 cDNA diluted 15 are different samples and. Lanes 2, 4, 6, and 8 are the respective cDNA<br />

diluted 150. Lane 9 is a PCR-negative control.<br />

3.2. cDNA Synthesis and Further Treatment<br />

1. A single reverse transcription with any of the two RT primers shown in Table 1 (G7AU2dT<br />

and GTGAU2dT) produced cDNA for a complete set of AU-DD reactions with matching<br />

PCR primers. These two different RT primers are anchored oligo(dT) 15 primers with a 5′ tag<br />

that accommodates the sequence of any of the AU-DD primers that will be used at the PCR<br />

step. The tag defines the particular RT primer and the set of AU-DD primers to be used. The<br />

15 T stretch was found to anneal more specifically on poly-A tails at the conditions used<br />

than others of 25 or longer. The tag was designed with the intention of generating the<br />

weaker secondary structure as possible. Standard oligo(dT) 15 was used to retrotranscribe<br />

control RNA samples (Jurkat and U937). First-stranded cDNA was prepared with<br />

Superscript II (Gibco-BRL) following manufacturer instructions with minor modifications<br />

indicated below.<br />

2. DNase treated total RNA (2–3 µg) in water and 10 pmoles of the chosen RT primer (Table 1)<br />

for a reaction volume of 20 µL were denatured at 72°C for 3 min on a PCR machine and<br />

chilled on ice for 1 min (see Note 10).<br />

3. The other reagents for a reaction volume of 20 µL except the enzyme are supplemented<br />

at room temperature (see Note 11). Annealing was allowed to proceed for 10 min at<br />

room temperature.<br />

4. 200 U SuperScript II (Gibco-BRL) were then added and the mixture was incubated for<br />

1 h at 42°C.<br />

5. Reverse transcription was stopped by heating at 90°C for 2 min and RNAse H was used as<br />

recommended by its supplier (Gibco-BRL). 1.8 U RNase H, 20 min at 37°C.<br />

6. To validate cDNA synthesis, normalization was performed by amplifying 110 and<br />

1500 cDNA dilutions for GAPDH in a 25-cycle PCR adjusted to an annealing temperature<br />

of 60°C (primers in Table 1B). Product concentration was estimated by visual inspection<br />

of ethidium bromide stained gels, and cDNAs were normalized by dilution according<br />

to their GAPDH equivalents if required. Typically no adjustment was needed (Fig. 2,<br />

see Note 12).<br />

7. Free nucleotides and primer were washed out in Qiaquick columns (Qiagen). cDNA was<br />

eluted in 50 µL of EE. The eluates were used directly in AU-DD.


AU-Differential Display 231<br />

Fig. 3. Typical AU-DD gel. Results from adherent monocytes by using a single PCR (not<br />

nested) are shown. Lanes 1 and 2 show fingerprints from activated cells. Lanes 3 and 4 show<br />

resting cells. Net cDNA was used for lanes 1 and 3 and diluted (15) for 2 and 4. Arrows signal<br />

fragments from differentially expressed genes.<br />

3.3. AU-DD<br />

At this stage, there are two desalted cDNA samples ready to be compared, the<br />

experimental and the control. Every reaction on a template is made in duplicate to<br />

detect variability caused by the sample concentration only. In this way, two cDNA<br />

concentrations, net and 15, are used in independent PCR tubes with otherwise identical<br />

primer and conditions (see Note 13). Reactions are prepared for 10-µL volumes and<br />

contain 2 µL of cDNA, primer at 3 µM (1.5 µM in nested reactions as proposed in<br />

Note 13), 10 mM Tris-HCl, pH 8.8; 1.5 mM MgCl 2 , 50 mM KCl, 0.1% Triton X-100,<br />

dNTPs each at 0.2 mM (see Note 14), 0.25 µL of α- 32 P-dATP and 50 mU/µL Platinum<br />

DNA polymerase (see Note 15). Reactions are run in a PTC-100 thermocycler (MJ<br />

Research) for 40 cycles with the following profile: 94°C for 40 s; 42°C for 1 min,<br />

20 s; and 70°C for 40 s.


232 Dominguez et al.<br />

3.4. Gel Electrophoresis<br />

1. AU-DD products (2 µL) are mixed with nondenaturing DNA loading buffer and separated<br />

in native 0.8-mm thick, 6% TBE polyacrylamide gels at 12 V/cm. Voltage was chosen<br />

such that it did not generate higher temperature than 45°C during the run to avoid heat<br />

denaturation. Every set of four reactions coming from the same cell line, both resting and<br />

activated, and at each of the two cDNA concentrations are run in adjacent lanes.<br />

2. Gels are fixed in 20% methanol10% acetic acid for 30 min before drying under vacuum<br />

at 80°C about 1 h with a Whatman 3 MM as support and autoradiographed as for standard<br />

sequencing gels (13, see Note 16).<br />

3. Individual bands (Fig. 3) are selected when unique or more intense in activated cells<br />

(see Note 17).<br />

4. The autorad is clamped against the dried gel by superimposing background signals with<br />

corresponding well line and gel edges. Edges of selected bands are punctured through the<br />

autorad with a hypodermic needle, leaving a mark in the dried gel. Gel slices containing<br />

those bands are cut out of the dried gel with a sterile scalpel knife.<br />

5. Dried gel pieces are rehydrated in 0.2 mL of EE for 3 h at 50°C. A further 1:20 dilution in<br />

EE was used as template for reamplification by 40 cycles of PCR and subsequent cloning.<br />

3.5. Confirmation of Differential Expression<br />

To confirm the differential expression, a semiquantitative RT-PCR was performed<br />

(Fig. 4, see Subheading 2.3., step 6). Specific PCR primers to the selected fragments<br />

were derived from the fragment sequences and were used in 40 cycles of PCR with<br />

paired resting and stimulated cDNAs. These cDNAs are normalized as described above,<br />

under Subheading 2.2., step 5.<br />

4. Designing a Primer Directed to AU Motifs<br />

A variety of AU-DD primers directed to single AU motifs were tested in preliminary<br />

experiments (not reported here). Their poor specificity and performance led us to design<br />

oligos with double AU pentamer motifs, which are often found in rich AU containing<br />

3′UTR of tightly controlled genes.<br />

AU-DD features reside in those specially designed primers that were used from<br />

reverse transcription to PCR. Results from experiments using two primers which differ<br />

in their 5′ anchors are presented: GTGAU2 and G7AU2 (Table 1). The 5′ anchors<br />

influence the annealing to natural AU containing sites and, therefore, the array of<br />

cDNAs that are selected and amplified during the PCR. In this way AU2 primers with<br />

different 5′ anchors sample distinct subsets of genes.<br />

The correctness of the rationale on the primer design was demonstrated in experiments<br />

which used an interleukin (IL)-2 gene fragment containing the AU2 sequence as<br />

template. In otherwise similar conditions the GTGAU2 primer allowed the detection of<br />

10 3 times less molecules of IL-2 cDNA than G7AU2 (not shown). A better anchoring<br />

of GTGAU2 on IL-2 does not necessarily imply that it also anneals better on other<br />

cDNAs. Therefore, both primers were used in parallel experiments, as other anchors<br />

can be tested, to increase the number of genes sampled.<br />

To keep structural simplicity, anchor domains containing only G or G and T were used,<br />

but variations can be expected to work as well as long as primer Tm is maintained.<br />

In an attempt to increase the affinity of the anchor, we tested inosine containing<br />

primers in different sequence configurations. When used alone, these primers reduced


AU-Differential Display 233<br />

Fig. 4. Confirmation of the differential expression by specific RT-PCR (primers derived from<br />

sequences of cloned AU-DD products) in agarose gels of a few AU-DD fragments taken as<br />

examples. Odd lanes are from activated cells and even lanes are the counterpart resting cells.<br />

the system sensitivity and reproducibility on the IL-2 cDNA. However, when used to<br />

supplement standard primers, a small but consistent gain in the number of products<br />

obtained from total cDNA was noticed, even when sensitivity to the IL-2 standard was<br />

not improved (not shown). Profile reproducibility among replicas was excellent for the<br />

primers and conditions described.<br />

5. Notes<br />

1. Different RNA preparation methods were examined in function of copurifying RNase<br />

activities, contaminating DNA, and yield. Ultracentrifugation in Cs salts (CsCl, CsTFA)<br />

gave an acceptable low level of DNA, but it was discarded because of its relatively<br />

poor yield. Phase lock heavy gel was found to improve both the yield in RNA and the<br />

partitioning of the genomic DNA to the lower phenolic phase. The single step acid-phenol<br />

method (14) was used with minor modifications intended to ensure RNA integrity and<br />

an efficient removal of both RNase activities and of contaminating genomic DNA. The<br />

procedure is described since this latter factor is considered relevant to the main method.<br />

2. Fresh tissue is snap frozen in liquid nitrogen and then stored at –80°C or lower until<br />

needed. Ceramic mortar is precooled by pouring on it liquid nitrogen where pestle is<br />

submerged; it is advised to wear gloves during this procedure. Initial bubbling will cease<br />

when the tool is cool enough. Then, the tissue is submerged and ground till reduced to<br />

a powdered state. The remaining liquid nitrogen is left to evaporate until the tissue is<br />

just wet, like a paste. Then, it is poured to a polypropylene 50-mL conical centrifuge<br />

tube. As soon as the tissue starts looking dry, add the lysis solution, mix by vortexing<br />

and homogenize.


234 Dominguez et al.<br />

3. If the container is not a polypropylene (PPN) tube, transfer lysates to a PPN centrifuge tube.<br />

Choose the tube so that the lysate volume is not larger than one third of its capacity.<br />

4. Because lower RNA amounts negatively affect the reliability of DD techniques, not less<br />

than 3 µg are processed.<br />

5. The RNA concentration in the storage ethanol solution is one fourth of the original<br />

concentration before adding the alcohol. In absence of salts, this solution forms a relatively<br />

uniform suspension of RNA, which is easy to pipet. To recover an aliquot, transfer the<br />

volume that contains the desired amount to a fresh tube. Supplement and mix with both<br />

coprecipitant (glycogen) and 0.1 vol of KAc, pH 5.0. Store an additional hour at –80°C<br />

and gather the sediment by centrifugation at 12,000g. After a single wash in 75% ethanol,<br />

the pellet is ready for any downstream application.<br />

6. To assure a complete removal of contaminating DNA, trial tests were performed. RNA<br />

from human thymus, which copurifies with relatively high content of DNA, was used for<br />

these assays. The best conditions were found when RNA (substrate) and DNase (enzyme)<br />

were kept as described.<br />

7. The absence of residual contaminating DNA was demonstrated by the failure to amplify<br />

the genomic locus of interferon-α, a multicopy gene, by PCR. A program of 40 cycles with<br />

an annealing temperature of 60°C was performed, with primers at 1 µM and reaction in<br />

10 µL (primers in Table 1). The inclusion of negative controls to test for PCR contamination<br />

is advisable. If DNA is detected, the DNAse-treated RNA sample is discarded.<br />

8. After DNase treatment, ethanol precipitation is apparently enough to inactivate any<br />

significant DNase activity without the need of any further treatment such as phenol.<br />

9. The RNA pellet can be stored indefinitely in this wash solution of 75% ethanol. It is<br />

left there before reverse transcription until it is confirmed by PCR that no residual DNA<br />

contamination has survived (see Note 7).<br />

10. Total RNA was always used because poly A selection from low amounts (tens of micrograms)<br />

of RNA was found inappropriate. The overall losses were significant, and an<br />

unquantifiable amount made the DD unreproducible and difficult to normalize.<br />

11. Given the 3′ location of the AU motifs there was a special interest in retrotranscribing true<br />

3′ ends. The separate step of annealing at room temperature was found to reduce primer<br />

extensions from false poly A tails.<br />

12. In this setting, the 1500 dilution always gives comparable results among samples; 4 µL<br />

checked by electrophoresis usually contains 5 to 10 ng of GAPDH amplimer.<br />

13. Because it is of interest to sample as many genes as possible different conditions are used<br />

side by side to increase throughput. Limited sample availability can also benefit from<br />

these variations. Primers of the series G7AU2 (Table 1) are used on cDNAs initiated<br />

from G7AU2dT. Similarly, GTGAU2 primers are more efficient on cDNA primed from<br />

GTGAU2dT. Both types of primers give different profiles, and different designs of the<br />

nonspecific 5′ domain also give different patterns. Proposed variations of conditions for a<br />

given cDNA to yield different profiles are as follows: (1) use the five primers of any series<br />

in 40 cycles of a single round of PCR; (2) run first a nonradioactive PCR of 30 cycles with<br />

either G7AU2 or GTGAU2, then dilute products 150 and run a series of nested PCR of<br />

30 cycles with the remaining four inner primers (G7AU2N or GTGAU2N) in four separate<br />

radioactive reactions; (3) reduce the concentration of cold dNTPs; (4) change the reaction<br />

buffer (different profiles can be achieved just by changing the Mg concentration in 1 mM<br />

differences); (5) combine any of these conditions.<br />

14. A cold dNTP concentration of 0.2 mM on radioactive reactions gives stronger signals and<br />

lower background than lower concentrations.


AU-Differential Display 235<br />

15. To reduce artifacts, it is important to use some hot start techniques or a preblocked<br />

polymerase. Some brands offer this kind of product. Those products that require extra<br />

PCR cycles for activation have been avoided. Note that different polymerases can also<br />

give different band patterns and sensitivity.<br />

16. Although even when fixed and washed radioactive gels still tend to give background<br />

signals on the autorad that assists the orientation, it may be preferred to use luminescent<br />

labels (such as the Glogos labels from Stratagene) for this purpose.<br />

17. According to the scientist’s interests, genes that are downregulated in experimental samples<br />

might also be of major interest. They would be isolated from the control sample since its<br />

concentration there would be higher.<br />

Acknowledgments<br />

Pepi Caro is warmly thanked for her expert technical support and patience during<br />

the development of this work.<br />

This work has been supported in part by the CDTI as Concerted Projects 95/0184<br />

and 97/0313.<br />

References<br />

1. Xu, N., Chen, C. Y., and Shyu, A. B. (1997) Modulation of the fate of cytoplasmic mRNA<br />

by AU-rich elements: key sequence features controlling mRNA deadenylation and decay.<br />

Mol. Cell <strong>Bio</strong>l. 17, 4611– 4621.<br />

2. Shaw, G. and Kamen, R. (1986) A conserved AU sequence from the 3′ untranslated region<br />

of GM-CSF mRNA mediates selective mRNA degradation. Cell 46, 659–667.<br />

3. Caput, D., Beutler, B., Hartog, K., Thayer, R., Brown-Shimer, S., and Cerami, A.<br />

(1986) Identification of a common nucleotide sequence in the 3’-untranslated region of<br />

mRNA molecules specifying inflammatory mediators. Proc. Natl. Acad. Sci. USA 83,<br />

1670–1674.<br />

4. Balmer, L. A., Beveridge, D. J., Jazayeri, J. A., Thomson, A. M., Walker, C. E., and<br />

Leedman, P. J. (2001) Identification of a novel AU-Rich element in the 3′ untranslated<br />

region of epidermal growth factor receptor mRNA that is the target for regulated RNAbinding<br />

proteins. Mol. Cell <strong>Bio</strong>l. 21, 2070–2084.<br />

5. King, L. M. and Francomano, C. A. (2001) Characterization of a human gene encoding<br />

nucleosomal binding protein nsbp1. Genomics 71, 163–173.<br />

6. Liang, P. and Pardee, A. B. (1992) Differential display of eukaryotic messenger RNA by<br />

means of the polymerase chain reaction. Science 257, 967–971.<br />

7. Welsh, J., Chada, K., Dalal, S. S., Cheng, R., Ralph, D., and McClelland, M. (1992)<br />

Arbitrarily primed PCR fingerprinting of RNA. Nucleic Acids Res. 20, 4965– 4970.<br />

8. Stone, B. and Wharton, W. (1994) Targeted RNA fingerprinting: the cloning of differentiallyexpressed<br />

cDNA fragments enriched for members of the zinc finger gene family. Nucleic<br />

Acids Res. 22, 2612–2618.<br />

9. Donohue, P. J., Alberts, G. F., Guo, Y., and Winkles, J. A. (1995) Identification by targeted<br />

differential display of an immediate early gene encoding a putative Serine/Threonine<br />

kinase. J. <strong>Bio</strong>l. Chem. 270, 10,351–10,357.<br />

10. Fischer, A., Saedler, H., and Theissen, G. (1995) Restriction fragment length polymorphismcoupled<br />

domain directed differential display: a highly efficient technique for expression<br />

analysis of multigene families. Proc. Natl. Acad. Sci. USA 92, 5331–5335.


236 Dominguez et al.<br />

11. Joshi, C. P., Kumar, S., and Nguyen, H. T. (1996) Application of modified differential<br />

display technique for cloning and sequencing of the 3′ region from three putative members<br />

of wheat HSP70 gene family. Plant Mol. <strong>Bio</strong>l. 30, 641–646.<br />

12. Dominguez, O., Ashhab, Y., Sabater, L., Belloso, E., Caro, P., and Pujol-Borrell, R. (1998)<br />

Cloning of ARE-containing genes by AU-motif-directed display. Genomics 54, 278–286.<br />

13. Sambrook, J., Fritsch, E. F., and Maniatis, T. (1989) Molecular Cloning: A Laboratory<br />

Manual. Cold Spring Harbor Laboratory Press, Plainview, NY.<br />

14. Chomczynski, P. and Sacchi, N. (1987) Single-step method of RNA isolation by acid<br />

guanidinium thiocyanate–phenol–chloroform extraction. Anal. <strong>Bio</strong>chem. 162, 156–159.


PCR Fluorescence Differential Display 237<br />

37<br />

PCR Fluorescence Differential Display<br />

Kostya Khalturin, Sergej Kuznetsov, and Thomas C. G. Bosch<br />

1. Introduction<br />

Differential display of mRNA via polymerase chain reaction (DD-PCR) has become<br />

a powerful procedure for the quantitative detection of differentially expressed genes<br />

in distinct cell populations (1–4).<br />

The standard procedure includes selective reverse transcription of polyadenylated<br />

RNA using specific anchored oligo(dT) primers, PCR amplification of cDNA using<br />

the oligo(dT) primer and an arbitrary upstream primer, resolution of PCR products on<br />

denaturing sequencing gels, and radioactive detection methods. To avoid hazardous<br />

radioisotopes, several nonradioactive methods for identification of differential display<br />

cDNAs have been reported, including ethidium bromide visualization in agarose gels<br />

(5), silver staining (6,7) and chemiluminescent detection (8,9) of cDNA bands.<br />

Several protocols have been published reporting the use of automated DNA sequencers<br />

for differential display of cDNAs labeled with infrared dyes (10) or fluorescent<br />

tags (11–13). A major drawback for using automated sequencers for this purpose<br />

is the inability to recover the amplified cDNAs from the gel. As an alternative, the<br />

programmable GenomyxLR DNA sequencer (Genomyx, Foster City, CA), which<br />

allows high resolution of cDNAs and easy localization and excision of radiolabeled<br />

bands from the gel, is becoming widely used in differential display studies (3).<br />

For nonradioactive detection of differential cDNA bands on the GenomyxLR DNA<br />

sequencer, cDNAs can be fluorescently labeled by using tetramethylrhodamine<br />

(TAMRA)-anchored primers. Alternatively, cDNAs can be fluorescently labeled<br />

by using TAMRA-dUTP (Perkin–Elmer), which is incorporated into extending<br />

cDNA during the PCR (14). Fluorescently labeled PCR products are separated on a<br />

GenomyxLR DNA sequencer (Genomyx), detected by the GenomyxSC Fluorescent<br />

Imaging Scanner, and directly excised from the gel for further characterization using an<br />

actual and virtual grid system. The flowchart shown in Fig. 1 outlines the experimental<br />

procedure.<br />

2. Materials<br />

1. peqGold RNAPure kit (PEQLAB <strong>Bio</strong>technologie GmbH, Germany) or other reagents<br />

for total RNA/mRNA extraction.<br />

2. DEPC water.<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

237


238 Khalturin, Kuznetsov, and Bosch<br />

Fig. 1. Flowchart of the fluorescent differential display procedure.<br />

3. First-Strand cDNA synthesis Kit (Amersham Pharmacia <strong>Bio</strong>tech Inc.).<br />

4. Tailing primers T 12 2NN (e.g., T 12 CA) at concentration 25 µM (see Note 2).<br />

5. Arbitrary 10-mer primers (e.g., OPA Kit by Operon Technologies Ltd., Amelada, CA)<br />

at concentrations of 5 µM.<br />

6. dNTPs stock solution (1 mM of each).<br />

7. Taq DNA Polymerase and 10, PCR buffer (Amersham Pharmacia <strong>Bio</strong>tech Inc.).<br />

8. TAMRA-dUTP (400 µM stock solution, Perkin–Elmer).<br />

9. MgCl 2 (25 mM stock solution).<br />

10. GeneQuantII photometer (Pharmacia) or equivalent device for measuring RNA<br />

concentration.


PCR Fluorescence Differential Display 239<br />

11. Thermal cycler with a ramping speed not more then 2.5°C/s. (or with programmable<br />

ramping speed).<br />

12. GenomyxLR DNA sequencer and GenomyxSC Fluorescent Imaging Scanner (Genomyx,<br />

Foster City, CA).<br />

3. Methods<br />

3.1. RNA Isolation (see Note 1)<br />

1. Isolate total RNA using peqGold RNAPureKit from approx 1 × 10 6 cells according to the<br />

manufacturer’s protocol. The final volume of total RNA suspension should be 20 µL.<br />

2. Measure OD 260 of 2 µL of RNA suspension in 80 µL of DEPC water (140 dilution). The<br />

total yield should be approx 300 µg of total RNA.<br />

3.2. Reverse Transcription (see Note 2)<br />

1. Reverse transcribe 3 µg of total RNA using First-Strand cDNA synthesis Kit.<br />

2. Place 3 µg of total RNA in a microcentrifuge tube and add RNAse-free (DEPC) water, if<br />

necessary, to bring the suspension volume to 7.75 µL.<br />

3. Heat the mixture for 10 min at 70°C and then place on ice immediately.<br />

Components (concentration of stock solution) Amount Final concentration<br />

Bulk First-Strand cDNA Reaction Mix (3×) 15 µL 1×<br />

DTT solution (200 mM ) 11 µL 13.3 mM<br />

Tailing primer, e.g. 5′-T 12 CA-3′ (25 µM ) 11.25 µL 2 µM<br />

RNA suspension (3 µg) 17.75 µL 0.2 µg/µL<br />

Final volume 15 µL<br />

4. Mix the following components on ice according to Table 1 for one reaction.<br />

5. Incubate for 1 h at 37°C.<br />

6. Incubate the tube at 95°C for 2 min to inactivate reverse transcriptase.<br />

7. Dilute the reaction mix 125. This quantity of template is sufficient for approx 180<br />

DD-PCRs in the volume of 10 µL.<br />

3.3. PCR (see Note 3)<br />

1. Set up the PCR mix according to Table 2.<br />

Components of stock solution) Amount Final concentration<br />

cDNA (125 dilution of RT-reaction mix) 12 µL –<br />

10× buffer 11 µL 111×<br />

MgCl 2 (25 mM ) 10.625 µL 113 mM<br />

random 10-mer primer (5 µM ) 10.5 µL 110.5 µM<br />

tailing primer (25 µM ) 11 µL 112.5 µM<br />

dNTP (1 mM of each) 12 µL 200 µM<br />

TAMRA-dUTP (400 µM ) 10.1 µL 114 µM<br />

Taq DNA Polymerase (5 U/µL) 10.1 µL 110.5 U<br />

H 2 O 12.675 µL –<br />

Final volume 10 µL<br />

2. Prepare all the samples in duplicates to check the reproducibility of the DD-PCR banding<br />

pattern.<br />

3. Perform the PCR with the following settings: initial denaturing 2 min at 95°C; 40 cycles<br />

30 s at 94°C (denaturation), 30 s at 40°C (annealing), and 30 s at 72°C (extension).<br />

4. Use the thermal cycler with a ramping speed not more then 2.5°C/s (see Note 4).


240 Khalturin, Kuznetsov, and Bosch<br />

3.4. Electrophoresis<br />

1. Before electrophoresis, add 7 µL of sample buffer (95% formamide, 0.25% dextran blue,<br />

10 mM EDTA) to each 10-µL sample.<br />

2. Denature amplified labeled fragments at 95°C for 3 min and load onto a high-resolution<br />

denaturing polyacrylamide gel (HR-1000 4.5% matrix, Genomyx, Foster City, CA).<br />

3. Perform the electrophoresis for 2.5 h in parallel using the GenomyxLR DNA sequencer<br />

(according to the manufacturer’s instructions).<br />

3.5. Detection of PCR Products, Elution, and Cloning<br />

1. cDNA bands are detected using the GenomyxSC Fluorescent Imaging Scanner (Genomyx).<br />

A typical gel with several differentially expressed transcripts of the freshwater polyp<br />

Hydra vulgaris is shown in Fig. 2 (filled and open arrows). Fluorescently labeled cDNA<br />

fragments usually range from about 100 to 2000 bp.<br />

2. For further characterization, DNA from differentially expressed bands is localized using<br />

the grid coordinate system provided with the GenomyxSC Fluorescent Imaging Scanner.<br />

After excision from the gel, cDNA in gel slices can easily be recovered and reamplified<br />

using described procedures (7,14).<br />

4. Notes<br />

In our minds, the most crucial steps for performing a successful differential display<br />

screening are as follows: (1) quality of the initial RNA taken for the cDNA preparation;<br />

(2) the quantity of cDNA used for the DD-PCR; and (3) the PCR cycling profile, which<br />

depends greatly on the type of thermal cycler used.<br />

1. In our experience DD-PCR works equally well in using either total RNA or mRNA. Kits<br />

for the isolation of poly(A) + mRNA, however, seem to be more reliable. The reason is that<br />

when total RNA is extracted, DNAse I treatment is usually needed to eliminate the DNA.<br />

However, this procedure sometimes can lead to considerable loss of mRNA.<br />

If extraction of total RNA is the method of choice, it is crucial to check the quality of the<br />

RNA obtained not only by the OD 260 and OD 280 measurement but also by electrophoresis<br />

of the RNA in an agarose gel. In the ethidium bromide-stained gel, the total RNA should<br />

appear as a smear of approx 500 to 2000 bp with two prominent bands of 28S rRNA and<br />

18S rRNA. If any RNA degradation or traces of high molecular weight DNA are observed,<br />

the RNA samples should not be used in DD-PCR.<br />

2. We prefer to synthesize cDNA using T (12) NN tailing primers (where NN stands for one of<br />

the 12 possible combinations of 4 nucleotides) instead of standard NotI primer. Using the<br />

same tailing primer both in reverse transcription and DD-PCR gives more distinct bands in<br />

comparison with the case when cDNA is made with NotI primer and then one of T (12) NN<br />

primers is used in a DD-PCR. A slight increase in the reaction temperature (40°C instead<br />

of the usual 37°C) during reverse transcription improves the specificity of cDNA synthesis.<br />

DD-PCR banding patterns obtained with the same random primer, but two different tailing<br />

primers (e.g., –5′-T 12 CA-3′ and 5′-T 12 AC-3′) should be considerably different.<br />

3. High final concentrations of dNTP (from 25 to 250 µM of each) and Mg 2+ (from 1.5 to<br />

4 mM) may help increasing the quantity and quality of bands. For optimal results, it is<br />

always necessary to try different dilutions of the cDNA sample (e.g., 125 → 150). A<br />

low final concentration of cDNA in DD-PCR can lead to nonreproducible banding patterns<br />

and even to differences within pares of tubes containing identical cDNA templates. One<br />

typical problem encountered when doing DD-PCR is that the tailing primer does not<br />

anneal during PCR, resulting in fragments flanked by random primer only. In that case,


PCR Fluorescence Differential Display 241<br />

Fig. 2. Typical picture of DD-PCR of H. vulgaris cDNA using two different random primers<br />

(OPA-9 and OPA-3) and 5′-T (12) CA-3′ tailing primer. Fluorescently labeled PCR products were<br />

visualized by GenomyxSC Fluorescent Imaging Scanner. Filled arrows, induced transcripts;<br />

open arrows, repressed transcripts; C, control animals; I, immuno challenged animals; *,<br />

unincorporated TAMRA-dUTP.


242 Khalturin, Kuznetsov, and Bosch<br />

the annealing temperature should be decreased from 40°C down to the 35°C. Besides that,<br />

the ratio between random 10-mer primer and tailing primer should be adjusted. Using<br />

fluorescent TAMRA-dUTP in the reaction mix adds additional complexity to the DD-PCR<br />

approach because it can be destroyed by frequent freeze-thaw cycles. PCR products<br />

labeled with new TAMRA-dUTP are nearly 10 times brighter than those labeled using<br />

TAMRA-dUTP that had undergone 3 to 4 freeze-thaw cycles. Because of that, TAMRAdUTP<br />

should be always kept in 2- to 3-µL aliquots. Usually, it is difficult to estimate<br />

the optimal concentration of TAMRA-dUTP for effective labeling. Therefore, a different<br />

dTTP:TAMRA-dUTP ratio should be tried (for example 301, 1001, 2001). Unincorporated<br />

TAMRA-dUTP results in the fluorescent cloud in the gel indicated by the asterisk<br />

in Fig. 2.<br />

4. One of the major variables during the DD-PCR experiment appears to be the fact that<br />

thermal cyclers vary greatly in their ramping time and other technical details. In our<br />

laboratory, we use two types of PCR machines: Omn-E (Hybaid) and RoboCycler<br />

(Stratagene). The former has a Peltier-element with the ramping time of approx 2.5°C/s<br />

(while cooling from 94° to 70°C), the latter nearly no ramping since a robotic arm transfers<br />

the samples from one thermoblock to another. When examining the influence of the<br />

ramping time on the DD-PCR and using the same temperature and time profiles in both<br />

machines (as well as identical components of the PCR mix) we observed considerably more<br />

fragments produced using the Omn-E thermal cycler in comparison with the Robocycler.<br />

The difference is not caused by the destruction of TAMRA-dye because the same result is<br />

observed in silver stained gels (7) as well as on those viewed by the fluorescent scanner.<br />

It seems the rapid change of the temperature provided by the RoboCycler prevents the<br />

efficient synthesis of PCR products when using standard PCR protocol. An increase of the<br />

annealing time from 30 to 90 s may overcome that drawback.<br />

Acknowledgments<br />

The work in the laboratory is supported by grants from the DFG (to T.C.G.B.) and<br />

by the Daimler-Benz Foundation (S.K.)<br />

References<br />

1. Liang, P., Averboukh, L., Keyomarsi, K., Sager, R., and Pardee, A. B. (1992) Differential<br />

display and cloning of messenger RNAs from human breast cancer versus mammary<br />

epithelial cells. Cancer Res. 52, 6966–6968.<br />

2. Liang, P. and Pardee, A. B. (1992) Differential display of eukaryotic messenger RNA by<br />

means of the polymerase chain reaction. Science 257, 967–971.<br />

3. Martin, K. J., Kwan, C. P., O’Hare, M. J., Pardee, A. B., and Sager, R. (1998) Identification<br />

and verification of differential display cDNAs using gene-specific primers and hybridization<br />

arrays. <strong>Bio</strong>Techniques 24, 1018–1026.<br />

4. McClelland, M., Mathieu-Daude, F., and Welsh, J. (1995) RNA fingerprinting and differential<br />

display using arbitrarily primed PCR (review). Trends Genet. 11, 242–246.<br />

5. Rompf, R. and Kahl, G. (1997) mRNA differential display in agarose gels. <strong>Bio</strong>Techniques<br />

23, 28–32.<br />

6. Bosch, T. C. G. and Lohmann, J. (1996) Non-radioactive differential display of messenger<br />

RNA, in Fingerprinting Methods Based on Arbitrarily Primed PCR, Springer Lab Manual<br />

(Micheli, M.R. and Bova, R., eds.), Springer Verlag, Heidelberg.<br />

7. Lohmann, J., Schickle, H. P., and Bosch, T. C. G. (1995) REN, a rapid and efficient<br />

method for non-radioactive differential display and isolation of mRNA. <strong>Bio</strong>Techniques<br />

18, 200–202.


PCR Fluorescence Differential Display 243<br />

8. An, G., Luo, G., Veltri, R. W., and O’Hara, S. M. (1996) Sensitive, nonradioactive differential<br />

display method using chemiluminescent detection. <strong>Bio</strong>Techniques 20, 342–346.<br />

9. Ross, R., Ross, X.-I., Rueger, B., Laengin, T., and Reske-Kunz, A. B. (1999) Nonradioactive<br />

detection of differentially expressed genes using complex RNA or DNA hybridization<br />

probes. <strong>Bio</strong>Techniques 26, 150–155.<br />

10. Motlik, J., Carnwath, J. W., Hermann, D., Terletski, V., Anger, M., and Niemann, H.<br />

(1998) Automated recording of RNA differential display patterns from pig granulosa cells.<br />

<strong>Bio</strong>Techniques 24, 148–153.<br />

11. Bauer, D., Müller, H., Reich, J., Riedel, H., Ahrenkiel, V., Warthoe, P., et al. (1993)<br />

Identification of differentially expressed mRNA species by an improved display technique<br />

(DDRT-PCR). Nucleic Acid Res. 21, 4272–4280.<br />

12. Jones, S. W. (1997) Generation of multiple mRNA fingerprints using fluorescence-based<br />

differential display and an automated DNA sequencer. <strong>Bio</strong>Techniques 22, 536–543.<br />

13. Luehrsen, K. R. (1997) Analysis of differential display RT-PCR products using fluorescent<br />

primers and Genescan software. <strong>Bio</strong>Techniques 22, 168–175.<br />

14. Reinhardt, B., Frank, U., Gellner, K., Lohmann, J., and Bosch, T. C. G. (1999) Highresolution,<br />

fluorescence-based differential display on a DNA sequencer followed by band<br />

excision. <strong>Bio</strong>Techniques 27, 268–271.


244 Khalturin, Kuznetsov, and Bosch


RNA Arbitrarily Primed PCR 245<br />

38<br />

Microarray Analysis Using RNA<br />

Arbitrarily Primed PCR<br />

Steven Ringquist, Gaelle Rondeau, Rosa-Ana Risques,<br />

Takuya Higashiyama, Yi-Peng Wang, Steffen Porwollik,<br />

David Boyle, Michael McClelland, and <strong>John</strong> Welsh<br />

1. Introduction<br />

RNA arbitrarily primed polymerase chain reaction (RAP-PCR) has been used<br />

extensively to identify differentially regulated genes (1–6). The RAP-PCR method<br />

begins with conversion of RNA into cDNA, followed by arbitrarily primed PCR.<br />

The technique uses arbitrarily primed PCR (7–11) to amplify cDNA stretches lying<br />

between sequences that, by chance, match arbitrarily chosen oligonucleotide primers<br />

well enough to initiate primer extension. In earlier applications of the method, the<br />

complex mixture of products was resolved by polyacrylamide gel electrophoresis<br />

(PAGE), yielding highly reproducible fingerprints characteristic of the RNA source.<br />

Differences between fingerprints resulting from differentially expressed genes were<br />

verified by Northern blot analysis or reverse transcription (RT)-PCR.<br />

Here, a combination of RAP-PCR and cDNA array technology is described (see<br />

Note 1, refs. 12–14), which provide dramatic improvement in detection sensitivity and<br />

ease of analysis when compared with PAGE. Typically, PAGE analysis of a RAP-PCR<br />

examines ~50 to 100 genes, whereas microarray analysis of RAP-PCRs examines<br />

~10 to 20% of the genes represented on the microarray, simultaneously. With PAGE,<br />

characterization of differentially regulated genes requires laborious purification of the<br />

band, cloning, and sequencing (15). With microarrays, the identity of the gene is known<br />

immediately because the arrayed sequences are known. As is generally understood with<br />

microarrays, the microarray-based RAP-PCR approach can quickly narrow all mRNAs<br />

down to a subset that exhibits regulation under the conditions examined. However, the<br />

RAP-PCR method selectively examines a different population of mRNAs than does<br />

probe from mRNA or total RNA: RAP-PCR selectively samples the complex, rare<br />

class of mRNA, presumably as the result of the greater likelihood of the arbitrary<br />

primers encountering a sufficiently good match in the complex class to enable primer<br />

extension. This enables the method to measure changes in RNA abundances for rare<br />

transcripts with greater ease than with simple oligo (dT) n priming of mRNA for probe<br />

synthesis. Iteration of the method using different arbitrary primers allows greater<br />

coverage of the mRNA population to be achieved.<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

245


246 Ringquist et al.<br />

In this chapter, protocols are presented for the synthesis of RAP-PCR probes for<br />

application to microarrays. Arrays spotted on nylon membranes can also be used.<br />

The manufacture of microarrays, themselves, is discussed only briefly, and the<br />

reader is reminded that there are a number of different approaches. For example,<br />

diverse chemistries for adhering cDNA sequences to glass slides, PCR-amplified vs.<br />

synthesized spotted sequences, different fluorescent dye-labeled nucleotides, different<br />

robotic spotting strategies, and other variables have been explored. The reader is<br />

encouraged to seek current technical <strong>info</strong>rmation in this rapidly moving field.<br />

2. Materials<br />

2.1. Special Reagents and Supplies<br />

1. RNeasy Mini Kit (Qiagen #74106, Valencia, CA).<br />

2. QIAshredder columns (Qiagen #79656, Valencia, CA).<br />

3. QIAquick PCR Purification Kit (Qiagen, #74106, Valencia, CA).<br />

4. Microcon YM-30 (Millipore #42410, Bedford MA).<br />

5. Human cDNA clones (I.M.A.G.E; Research Genetics, Huntsville, AL).<br />

6. Vector-specific primers (Genosys <strong>Bio</strong>technologies, The Woodlands, TX): Forward:<br />

5′-ctgcaaggcgattaagttgggtaac-3′, Reverse:5′-gtgagcggataacaatttcacacaggaaacagc-3′<br />

7. First strand cDNA primer: oligo(dT) 20 .<br />

8. Arbitrary primers: primer A: (5′-ACGAAGAAGAAGAG), primer B (5′-GTGACAGACA;<br />

Genosys <strong>Bio</strong>technologies, The Woodlands, TX).<br />

9. [α- 32 P] dCTP 10 Ci/mL (ICN, Irvine, CA).<br />

10. Random hexamers (NNNNNN; Genosys <strong>Bio</strong>technologies, The Woodlands, TX).<br />

11. Cy3-dCTP and Cy5-dCTP (Amersham, Piscataway, NJ).<br />

2.2. Cell Culture<br />

1. Fibroblast (ATCC #CRL 2091, Manassas, VA).<br />

2. Cell-culture dishes (150 cm; Nunc, Rochester, NY).<br />

3. Cell culture media (DMEM, Irvine Scientific #9031 Irvine, CA), plus 10% fetal bovine<br />

serum (Omega scientific #FB-01, Tarzana, CA).<br />

4. Penicillin (1000 U/mL).<br />

5. Streptomycin (1 mg/mL).<br />

2.3. Enzymes<br />

1. M-MLV reverse transcriptase (200 U/µL; Promega, Madison, WI).<br />

2. AmpliTaq DNA polymerase Stoffel fragment (10 U/µL; Perkin–Elmer Cetus, Norwalk,<br />

CT).<br />

3. RNase-free DNase (10 U/µL) (Roche Molecular <strong>Bio</strong>chemicals, Indianapolis, IN).<br />

4. RNase inhibitor (40 U/µL) (Roche Molecular <strong>Bio</strong>chemicals, Indianapolis, IN).<br />

2.4. Common Reagents, Supplies, Buffers, and Equipment<br />

1. DMSO.<br />

2. Distilled, deionized H 2 O.<br />

3. 1× TE (pH 8.0).<br />

4. Formamide dye solution.<br />

5. 4% polyacrylamide, 8 M urea sequencing-style gels, prepared with 1× TBE buffer.<br />

6. 1% agarose minigels (1× TBE).<br />

7. 5× DNase I digestion buffer: 100 mM Tris-HCl, 50 mM MgCl 2 , pH 8.0.


RNA Arbitrarily Primed PCR 247<br />

8. 5× reverse transcription mixture: 250 mM Tris-HCl, pH 8.3, 375 mM KCl, 15 mM MgCl 2 ,<br />

100 mM DTT, 1.0 mM each dNTP, 2.5 µM oligo(dT) 20 , and 100 U M-MLV reverse<br />

transcriptase.<br />

9. 2× RAP-PCR mixture: 20 mM Tris-HCl, pH 8.3, 20 mM KCl, 6 mM MgCl 2 , 0.35 mM each<br />

dNTP, 2 µM each arbitrary primer (see text); 2 µCi [α- 32 P] dCTP, and 0.5 U/µL AmpliTaq<br />

DNA polymerase Stoffel fragment.<br />

10. 2× Klenow reaction mixture: 20 mM Tris-HCl, pH 7.5, 10 mM MgCl 2 , 15 mM DTT,<br />

0.050 mM dATP, 0.050 mM TTP, 0.050 mM dGTP, 0.018 mM dCTP, 40 U Klenow.<br />

11. 2× terminal transferase buffer: 400 mM potassium cacodylate; 50 mM Tris-HCl, pH 7.2;<br />

500 mg/mL bovine serum albumin, and 3 mM CoCl 2 .<br />

12. Blocking solution: 10 µg/µL oligo (dA) 20 , 10 µg/µL yeast tRNA, 10 µg/µL human Cot-1<br />

DNA.<br />

13. 20× SSC.<br />

14. Human Cot-1 DNA (Invitrogen #15279, Carlsbad, CA).<br />

15. Klenow (New England <strong>Bio</strong>Labs #MO210S, Beverly, MA).<br />

16. 96-well microtiter plates.<br />

17. 96-well PCR amplification plates.<br />

18. Ultraviolet spectrophotometer.<br />

19. Oven for baking slides.<br />

20. Vacuum desiccator for slide storage.<br />

21. Kodak <strong>Bio</strong>Max X-Ray (Kodak #8715187, Rochester, NY).<br />

22. Succinic anhydride (Sigma #S-7626, St. Louis, MO).<br />

23. 1-methyl-2-pyrrolidinone (Sigma #6762, St. Louis, MO).<br />

24. poly-L-lysine (Sigma #P8920, St. Louis, MO).<br />

2.5. Special Equipment<br />

1. PCR amplification machine (GeneAmp ® PCR System 9700, Perkin-Elmer Cetus, Norwalk,<br />

CT).<br />

2. High-speed robotic arrayer (OmniGrid, GeneMachine, San Carlos, CA).<br />

3. Ultraviolet irradiation device (Stratagene, La Jolla, CA).<br />

4. ScanArray 5000 (GSI Lumonics, Billerica, MA).<br />

5. QuantArray analysis software (GSI Lumonics, Billerica, MA).<br />

6. GoldSeal glass slides (Fisher Scientific, Pittsburgh, PA).<br />

Molecular biology grade, RNase free reagents are used. Sterile disposable polypropylene<br />

was used rather than glassware.<br />

3. Methods<br />

3.1. Preparation of RNA<br />

1. Fibroblasts were grown in culture to about 7 × 10 6 cells per plate and harvested by scraping in<br />

the presence of RTL buffer lysis buffer (RNeasy kit) and homogenized through Qiashredder<br />

columns, both according to the manufacturer’s instructions. Total RNA was isolated using<br />

the RNeasy total RNA purification kit and eluted into 50 µL of distilled water.<br />

2. To 49 µL total RNA was added 6 µL of DNase I digestion buffer, 5 µL of DNase I, and<br />

0.5 µL of RNase inhibitor for 30 min at 37°C.<br />

3. DNase-treated total RNA was purified again using the RNeasy total RNA kit according to<br />

manufacturer’s instructions and eluted into 50 µL of distilled water.<br />

4. RNA concentration was determined by spectrophotometry in TE buffer (10 mM Tris-HCl,<br />

pH 8.4; 0.1 mM EDTA) assuming that one OD unit was equivalent to 40 µg/mL of RNA.


248 Ringquist et al.<br />

5. Total RNA samples were adjusted to 400 ng/µL with distilled water, checked for the<br />

integrity of the rRNAs by 1% agarose gel electrophoresis, and stored at –20°C.<br />

3.2. cDNA Synthesis<br />

1. Reverse transcription was performed on total RNA using at least three twofold dilutions of<br />

total RNA per sample between 1000 and 125 ng per 10 µL of reaction (see Note 1). cDNA<br />

synthesis uses oligo(dT) 20 for first strand priming, or alternatively, an arbitrary sequence<br />

primer can be used under exactly the same conditions. The reaction mixture contained<br />

20 µL of purified, DNase I-treated RNA plus 5 µL of 5× reverse transcription mixture.<br />

cDNA reactions were performed at room temperature (usually 23°C) for 15 min followed<br />

by incubation at 37°C for 1 h.<br />

2. The reactions were stopped by heating for 5 min in a boiling water bath followed by<br />

cooling on ice. Reactions should be diluted fourfold with 75 µL of distilled water and<br />

stored at –20°C, or one can proceed to the next step. (see Note 2).<br />

3.3. RAP-PCR<br />

1. 2× RAP-PCR mixture was prepared using different pairs of arbitrary primers (see Note 3).<br />

In Fig. 1, primers A and B were used.<br />

2. Diluted cDNAs (10 µL) were mixed with the same volume of 2X RAP-PCR mixture.<br />

Thermocycling was performed using an initial 3-min incubation at 94°C followed by 35<br />

cycles of 94°C for 1 min, 35°C for 1 min, and 72°C for 2 min.<br />

3.4. Polyacrylamide Gel Electrophoresis<br />

1. An aliquot of the amplification products (2 µL) was mixed with 9 µL of formamide dye<br />

solution, denatured at 85°C for 4 min, and chilled on ice. A sample of 2 µL was loaded<br />

onto a 4% polyacrylamide, 8 M urea gel, prepared with 1× TBE buffer. The PCR products<br />

resulting from different concentrations of the same RNA template were loaded side by<br />

side on the gel.<br />

2. Electrophoresis was performed at 1700 V or at a constant power of 50 to 70 W until the<br />

xylene cyanol tracking dye reached the bottom of the gel (about 4 h). The gel was dried<br />

under vacuum and either placed on Kodak <strong>Bio</strong>Max X-Ray film for 16 to 48 h or used to<br />

expose a phosphor screen (Molecular Dynamics, Sunnyvale, CA) and analyzed on a Storm<br />

820 (Molecular Dynamics, Sunnyvale, CA). The results are shown in Fig. 1.<br />

3.5. Fluorescent Labeling<br />

1. Up to 10 µg of PCR product from the RAP-PCR can be purified using a QIAquick PCR<br />

Purification Kit (Qiagen, Valencia, CA), which removes unincorporated bases, primers,<br />

and primer dimer less than 40 bp. DNA was recovered in 50 µL of 10 mM Tris-HCl<br />

(pH 8.3), usually yielding 2 to 3 µg of DNA per RAP-PCR amplification. Recovered PCR<br />

products were quantified by spectroscopy assuming 33 µg/mL/OD unit.<br />

2. Klenow labeling of RAP-PCR DNA was performed by random primed synthesis using up<br />

to 1 µg of PCR product per reaction. PCR product was mixed with 12 µg total of 6-mer or<br />

9-mer random sequence primer and boiled for 4 min, cooled to room temperature, and spun<br />

briefly in a microfuge. The volume was adjusted to 20 µL, 25 µL of 2× Klenow reaction<br />

mixture, and 5 µL of either 1 mM Cy3-dCTP or 1 mM Cy5-dCTP (Amersham, Piscataway,<br />

NJ) was added and the mixture was incubated at 37°C for 4 h. The reaction was stopped by<br />

incubation at 70°C for 10 min to inactivate the Klenow fragment. Alternatively, addition of<br />

EDTA pH 8 to a final concentration of 10 mM can also be used. Probes from oligo (dT) n can<br />

be prepared by standard means to examine more abundant transcripts (see Note 4).


RNA Arbitrarily Primed PCR 249<br />

Fig. 1. RAP-PCR fingerprints. Products from RAP-PCR were resolved by electrophoresis<br />

as described. DNA molecular weights were determined using 32 P-labeled DNA fragments<br />

from MspI-digested pBR322 plasmid DNA, lane 1. Lanes 2 through 25 contain RAP-PCR<br />

products prepared from human fibroblast total RNA. Fingerprints were prepared from fibroblasts<br />

synchronized by serum starvation for 48 h and then treated by addition of serum containing<br />

media for the indicated times prior to isolation of total RNA, including isolation of total RNA<br />

from nonserum-starved cells. Fingerprinting of each sample was performed using three different<br />

concentrations of total RNA in the cDNA reaction, that is, 500, 250, and 125 ng per 10 µL of<br />

reaction. The fourth (empty) lane in each sample is the reverse transcriptase minus control.


250 Ringquist et al.<br />

3. Purification of fluorescently labeled material from unincorporated dye was performed by<br />

centrifuging the sample in a microfuge using Microcon columns twice, using 500 µL TE<br />

(pH 8.0) washes each time, according to the manufacturer’s instructions.<br />

4. Labeling of PCR products can also be performed by Terminal Transferase-catalyzed<br />

addition of Cy3-dCTP or Cy5-dCTP, as well as other fluorescent-labeled or modified<br />

deoxynucleotide triphosphates. 1 µg of PCR product in 20 µL was combined with<br />

25 µL 2× terminal transferase buffer and 5 µL of 1 mM Cy5-dCTP or 1 mM Cy3-dCTP.<br />

The reaction was performed at 37°C for 1 h and purified using the Microcon columns<br />

as described previously.<br />

5. After purification over microcon columns, fluorescent probe was dried under vacuum and<br />

dissolved in 8 µL of 1× TE buffer. Preparation of probe in hybridization solution was<br />

performed by addition of 8 µL of probe solution to 2 µL of blocking solution, 2.1 µL of<br />

20× SSC, and 0.4 µL of 10% SDS. The solution was then incubated 1 min in a boiling<br />

water bath, and allowed to cool on the bench top for 30 min.<br />

3.6. Microarray Hybridization<br />

1. Preparation of microarrays followed the protocol of Eisen and Brown (16) and used<br />

polylysine-treated slides and PCR products generated from sequenced human cDNA<br />

libraries; please see Eisen and Brown (16) for details on preparing polylysine-coated slides.<br />

Inserts from the IMAGE library, supplied by Research Genetics, were PCR amplified by<br />

standard means using the vector-specific primers. Printing of PCR products was done<br />

using an Omnigrid robot to spot PCR products onto GoldSeal glass slides using a 16 pin<br />

print head. After printing, slides were rehydrated briefly over an 80°C waterbath and snap<br />

heated for about 6 s by placing the slides on an 80°C heat block. Ultraviolet treatment was<br />

performed using a Stratalinker 2400 (Stratagene, La Jolla, CA) at 60 mJ for 1 min. The<br />

slides were then baked at 80°C for 2 h and stored.<br />

2. Immediately before hybridization, slides were treated with succinic anhydride as described<br />

by Eisen and Brown (16). The succinic anhydride blocking solution must be prepared<br />

fresh by placing 6 g of succinic anhydride into a dry beaker followed by addition of<br />

335 mL of 1-methyl-2-pyrrolidinone. After the succinic anhydride was dissolved 15 mL<br />

of 1 M sodium borate, pH 8.0 was added. The solution was stirred for 5 s to mix the<br />

solutions. The mixture was then poured into a clean glass chamber and the slides placed<br />

quickly inside and incubated for 15 min.<br />

3. Immediately after succinic anhydride blocking, slides were placed in a chamber with<br />

boiling water (the volume should be at least twice that of the blocking solution). The slides<br />

were plunged up and down and shaken for 2 min, and then the slides were immediately<br />

transferred to a chamber containing 95% ethanol, agitated and then dried for 37°C for<br />

15 min.<br />

4. Hybridization was performed by placing one glass chamber with distilled water in a 65°C<br />

oven for at least 30 min before doing the hybridization. One hybridization chamber was<br />

cleaned with pressurized air, and 2 drops (10 µL) of 3XSSC (450 mM sodium chloride,<br />

45 mM sodium citrate, pH 7.0) were placed in the wells. The slides were then cleaned with<br />

pressurized air and placed in the hybridization chamber. The probe was then pipetted onto<br />

the slide, usually in 40 µL, taking care to avoid bubbles. The coverslip was then cleaned<br />

with pressurized air and placed over the probe. This volume of probe usually spreads easily<br />

along the surface; if not, the overslip can be pressed to help the spreading. The chamber<br />

was then sealed and submerged in a 65°C water bath for 4 to 16 h.<br />

5. The hybridization was stopped by washing in 2× SSC, 0.1% SDS, followed by washing<br />

in 1× SSC. Samples were incubated for 2 min with 0.2× SSC. The washing steps were<br />

performed at room temperature. Drying should be avoided in all the washing steps. Prepare


RNA Arbitrarily Primed PCR 251<br />

Fig. 2. Scatter plot of fluorescent intensity of RAP-PCR products analyzed by microarray<br />

hybridization. Samples were RAP-PCR products (1 µg) from duplicate fingerprint reactions labeled<br />

using Cy3-dCTP (green fluorescence) or Cy5-dCTP (red fluorescence), random primer and<br />

Klenow as described. The microarray contained 6400 double spotted cDNAs of known identity.<br />

400 mL of each solution and place them in three different glass chambers. To minimize<br />

the transfer of solutions, a different container should be used each time and only the slides<br />

should be transferred. The solution should then be spun down 5 min to dry the slides.<br />

3.7. Analysis of Microarray Data<br />

Analysis of the microarray hybridization was performed using a ScanArray 5000<br />

or equivalent instrument to monitor fluorescent signal. Data was gathered first from<br />

the Cy5 dye followed by scanning of the Cy3 fluorescent signal. Typical instrument<br />

settings were laser power at 50 and photomultiplier tube at 100. Microarrays were<br />

scanned using 10-µm resolution. Quantification of microarray data was performed<br />

using QuantArray software or an equivalent software package. The results are shown<br />

in a scatter plot in Fig. 2. More than 95% of the signals that are significant relative to<br />

negative controls from salmonella are within twofold.<br />

4. Notes<br />

1. Knowing the total RNA concentration is critical for RAP-PCR. cDNA is synthesized<br />

at several concentrations ranging between 1000 ng and 125 ng of total RNA per 10-µL<br />

reaction. At lower concentrations of input total RNA there sometimes arise spurious PCR<br />

products, which overwhelm the amplification step. Also, DNA contaminants are monitored<br />

by the inclusion of a reverse transcriptase-free control in initial RAP-PCR experiments.<br />

This reaction is set up exactly as a RAP-PCR, but reverse transcriptase is omitted and<br />

replaced with water. The result should be a gel lane with no bands.


252 Ringquist et al.<br />

2. The present method is based closely on the protocol described by Trenkle et al. (12,14),<br />

in which RAP-PCR was used to generate probes for differential screening of cDNA<br />

arrays on nylon membranes. Each array contained 18,432 cDNA clones from the IMAGE<br />

consortium and hybridization detected approx 1000 cDNA clones using each RAP-PCR<br />

probe. Different RAP-PCR fingerprints gave hybridization patterns having very little<br />

overlap (less than 3%) with each other or with hybridization patterns from total cDNA<br />

probes. Thus, repeated application of RAP-PCR probes allows a greater fraction of the<br />

message population to be screened on this type of array than can be achieved with a<br />

radiolabeled total cDNA probe.<br />

3. In general, there are no constraints on the primers except that they contain at least a few<br />

C or G residues, that the 3′-ends are not complementary with themselves or the other<br />

primer in the reaction, to avoid primer dimer, and that primer sets are chosen that are<br />

different in sequence so that the same parts of mRNA are not amplified in different<br />

fingerprints. PCR with a combination of arbitrary and oligo(dT) primers, as has been used<br />

in differential display (17–20), can also be used to generate an effective probe for cDNA<br />

arrays. The use of different oligo(dT) anchor primers with the same arbitrary primer results<br />

in considerable overlap among the genes sampled by each probe. This can be avoided by<br />

using different arbitrary primers with each oligo(dT) anchor primer (13).<br />

4. Total RNA as well as poly(A) purified mRNA can also be labeled during reverse transcription<br />

for microrray analysis. Methods for preparing fluorescent cDNA from RNA have been<br />

presented by Eisen and Brown (16).<br />

References<br />

1. Welsh, J., Chada, K., Dalal, S. S., Cheng, R., Ralph, D., and McClelland, M. (1992)<br />

Arbitrarily primed PCR fingerprinting of RNA. Nucleic Acids Res. 20, 4965– 4970.<br />

2. Vogt, T. M., Welsh, J., Stolz, W., Kullmann, F., Jung, B., Landthaler, M., et al. (1997)<br />

RNA fingerprinting displays UVB-specific disruption of transcriptional control in human<br />

melanocytes. Cancer Res. 57, 3554–3561.<br />

3. Vogt, T., Stolz, W., Welsh, J., Jung, B., Kerbel, R. S., Kobayashi, H., et al. Overexpression of<br />

Lerk-5/Eplg5 messenger RNA: A novel marker for increased tumorigenicity and metastatic<br />

potential in human malignant melanomas. Clin. Cancer Res. 4, 791–797.<br />

4. Ralph, D., McClelland, M., and Welsh, J. (1993) RNA fingerprinting using arbitrarily primed<br />

PCR identifies differentially regulated RNAs in mink lung (Mv1Lu) cells growth arrested<br />

by transforming growth factor beta 1. Proc. Natl. Acad. Sci. USA 90, 10,710–10,714.<br />

5. McClelland, M. and Welsh, J. (1994) RNA fingerprinting by arbitrarily primed PCR. PCR<br />

Methods Appl. 4, S66–S81.<br />

6. McClelland, M., Mathieu-Daude, F., and Welsh, J. (1995) RNA fingerprinting and differential<br />

display using arbitrarily primed PCR. Trends Genet. 11, 242–246.<br />

7. Welsh, J. and McClelland, M. (1991) Genomic fingerprinting using arbitrarily primed PCR<br />

and a matrix of pairwise combinations of primers. Nucleic Acids Res. 19, 5275–5279.<br />

8. Welsh, J., Petersen, C., and McClelland, M. (1991) Polymorphisms generated by arbitrarily<br />

primed PCR in the mouse: application to strain identification and genetic mapping. Nucleic<br />

Acids Res. 19, 303–306.<br />

9. Welsh, J. and McClelland, M. (1990) Fingerprinting genomes using PCR with arbitrary<br />

primers. Nucleic Acids Res. 18, 7213–7218.<br />

10. McClelland, M. and Welsh, J. (1994) DNA fingerprinting by arbitrarily primed PCR. PCR<br />

Methods Appl. 4, S59–S65.<br />

11. Williams, J. G., Kubelik, A. R., Livak, K. J., Rafalski, J. A., and Tingey, S. V. (1990)<br />

DNA polymorphisms amplified by arbitrary primers are useful as genetic markers. Nucleic<br />

Acids Res. 18, 6531–6535.


RNA Arbitrarily Primed PCR 253<br />

12. Trenkle, T., Welsh, J., Jung, B., Mathieu-Daude, F., and McClelland, M. (1998) Nonstoichiometric<br />

reduced complexity probes for cDNA arrays. Nucleic Acids Res. 26,<br />

3883–3891.<br />

13. Trenkle, T., Welsh, J., and McClelland, M. (1999) Differential display probes for cDNA<br />

arrays. <strong>Bio</strong>Techniques. 27, 554–560,562,564.<br />

14. Trenkle, T., Mathieu-Daude, F., Welsh, J., and McClelland, M. (1999) Reduced complexity<br />

probes for DNA arrays. Methods Enzymol. 303, 380–392.<br />

15. Mathieu-Daude, F., Cheng, R., Welsh, J., and McClelland, M. (1996) Screening of<br />

differentially amplified cDNA products from RNA arbitrarily primed PCR fingerprints<br />

using single strand conformation polymorphism (SSCP) gels. Nucleic Acids Res. 24,<br />

1504–1507.<br />

16. Eisen, M. B. and Brown, P. O. (1999) DNA arrays for analysis of gene expression. Methods<br />

Enzymol. 303, 179–205.<br />

17. Liang, P. and Pardee, A. B. (1992) Differential display of eukaryotic messenger RNA by<br />

means of the polymerase chain reaction. Science 257, 967–971.<br />

18. Liang, P., Zhu, W., Zhang, X., Guo, Z., O’Connell, R. P., Averboukh, L., et al. (1994)<br />

Differential display using one-base anchored oligo-dT primers. Nucleic Acids Res. 22,<br />

5763–5764.<br />

19. Liang, P. and Pardee, A. B. (1998) Differential display. A general protocol. Mol. <strong>Bio</strong>technol.<br />

10, 261–267.<br />

20. Martin, K. J. and Pardee, A. B. (1999) Principles of differential display. Methods Enzymol.<br />

303, 234–258.


254 Ringquist et al.


Arrays for Genotyping 255<br />

39<br />

Oligonucleotide Arrays for Genotyping<br />

Enzymatic Methods for Typing Single Nucleotide Polymorphisms<br />

and Short Tandem Repeats<br />

Stephen Case-Green, Clare Pritchard, and Edwin Southern<br />

1. Introduction<br />

Much of modern genetics is based on analysis of DNA sequence. Therefore, there<br />

is great pressure to scale up sequence analysis while decreasing its cost. The most<br />

promising platforms are based on the use of oligonucleotide arrays (DNA chips), which<br />

perform many analyses in parallel (1). Arrays comprise libraries of oligonucleotides<br />

or polynucleotides attached to solid supports at defined locations. Assays can be<br />

performed on the whole library simultaneously, increasing both the speed and accuracy<br />

of analysis of the target nucleic acid.<br />

In this chapter, we describe the fabrication and some uses of oligonucleotide arrays<br />

that are under development in our laboratory and give an idea of the flexibility of the<br />

array platform. Practical array-based assays will need to be individually optimized to<br />

the desired target nucleic acid. Although arrays can be made from large fragments of<br />

DNA (2), this chapter concentrates on arrays of synthetic oligonucleotides. Analytic<br />

methods for single nucleotide polymorphisms (SNPs) and short tandem repeat (STR)<br />

measurement are described. Other uses have been made of DNA arrays, such as<br />

expression monitoring (3) and antisense oligonucleotide optimization (4).<br />

Most methods for DNA sequence analysis are based on gel electrophoretic separation<br />

of DNA fragments. Automation and miniaturization of these techniques has been<br />

difficult to achieve. A major advantage of arrays is the potential for automation at all<br />

stages of manufacture and application. Figure 1 shows a scheme for an assay using<br />

an oligonucleotide array.<br />

1.1. Fabrication of Arrays<br />

Two methods of array fabrication are in general use, listed as follows:<br />

1. Presynthesized oligonucleotides can be applied to a support.<br />

2. Oligonucleotide synthesis can be performed in situ at specific sites on the support.<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

255


256 Case-Green, Pritchard, and Southern<br />

Fig. 1. General scheme for assays using oligonucleotide arrays.<br />

Both methods have advantages and disadvantages, as summarized here.<br />

• In situ fabrication allows a truly combinatorial synthesis of oligonucleotide probes to<br />

be performed, whereas presynthesis involves synthesis of every probe and its individual<br />

application to the array.<br />

• The combinatorial fabrication route may be best suited to producing low numbers or single<br />

copies of many different arrays, whereas presynthesis allows enough pure oligonucleotide<br />

to be produced to make many copies of the same array with speed and reliability.<br />

• Presynthesized oligonucleotides can be purified before addition to the array, whereas<br />

in situ fabrication depends on the reliability and high yields of conventional solid phase<br />

synthesis to ensure that most of the array probe products are correct.<br />

In our laboratory, we have developed barrier-based methods for combinatorial in<br />

situ array synthesis using standard oligonucleotide synthesis reagents (5) confined<br />

within reaction cells (Fig. 2).<br />

These techniques allow combinatorial synthesis of multiple oligonucleotides using<br />

four basic steps (6). Different shapes of cells are used to create arrays with different<br />

organizations and uses. For example, a square-shaped cell can be used for the synthesis<br />

of an array containing overlapping oligonucleotide sequences that have all the possible<br />

complements of a target sequence up to a chosen length of oligonucleotide.<br />

Alternatively, using a block with channels cut into it allows serial synthesis of<br />

oligonucleotides of any sequence in a series of stripes (7,8). If the oligonucleotides<br />

on an array are arranged in parallel lines, it may be used in more than one reaction<br />

if the target samples are hybridized in lines perpendicular to the oligonucleotides (8)<br />

or the array cut into strips.<br />

Assays described in this chapter are performed on amino functionalized polypropylene<br />

support (Beckman) (7,9). This surface is suitable as a support for solid-phase<br />

oligonucleotide synthesis. All syntheses in this chapter use 3′-dimethoxytrityl-5′-N,N′-<br />

diisopropylcyanoethylphosphoramidites (reverse phosphoramidites). Deprotection


Arrays for Genotyping 257<br />

Fig. 2. The layout of reaction cells used for the synthesis of arrays. The synthesis support<br />

is pressed to the reaction cell by a clamp. The inlet and outlet ports are attached to a DNA<br />

synthesizer. DNA synthesis on the support is achieved using the synthesizer to add and remove<br />

reagents from the reaction cell; the reagents flood the surface of the support. After the required<br />

sequence has been added the clamp is released and the cell moved and reclamped to allow<br />

further synthesis. Apparatus described in ref. 6 can be used to aid this process.<br />

leaves the oligonucleotides attached to the polypropylene via the initial, ammonia<br />

stable, phosphoramidite linkage to the 5′ end of the oligonucleotide and the amine<br />

of the polypropylene support.<br />

1.1.1. Arrays for Biallelic (SNP) Typing<br />

The simplest fabrication method involves the synthesis of each allele specific<br />

oligonucleotide (ASO) one at a time at a separate site on the surface. A patch of the<br />

first oligonucleotide is synthesized on the support using a cell to direct the synthesis<br />

reagents. The cell is then moved to a position adjacent to the first oligonucleotide and<br />

the second oligonucleotide synthesized. A long narrow cell (Fig. 2, left) minimizes


258 Case-Green, Pritchard, and Southern<br />

reagent use and allows the array to be cut into strips in a direction perpendicular to<br />

the array synthesis.<br />

Alternatively, a sequence that flanks the variable base is synthesized; the cell is<br />

displaced by half a cell’s width and a base corresponding to one of the alleles is added,<br />

the cell is then moved to cover the other half and the other varying base added. The<br />

cell can then be returned to its original position and any further bases common to both<br />

allele specific probe oligonucleotides added.<br />

1.1.2. STR Array Fabrication<br />

The basic requirements for an array suitable for use in STR typing are shown in<br />

Fig. 3. All the oligonucleotides of the array include a sequence complementary to the<br />

region immediately adjacent to the repeats. This registration sequence forms duplex<br />

between the target and the array oligonucleotides and aligns the start of the repeat<br />

of the oligonucleotides with that in the target. The array oligonucleotides vary in the<br />

number of repeat units that they contain on top of the flanking registration sequence.<br />

Synthesis of these arrays can be carried out in a combinatorial manner. Figure 4 shows<br />

a scheme for the synthesis of an array for typing the FES locus (10).<br />

1.2. Preparation of Target DNA<br />

To achieve good hybridization yields and allow efficient extension of the array<br />

oligonucleotides, a single-stranded target is required. Depending on the type of assay<br />

the target may be either labeled or unlabeled. Because the DNA for analysis is usually<br />

genomically derived material, an amplification step is normally performed.<br />

For hybridization assays, the most common target species is body labeled RNA,<br />

prepared by carrying out a polymerase chain reaction (PCR) of the required region<br />

using a primer that includes an RNA polymerase promoter sequence and transcribing<br />

the product with the appropriate enzyme. The reaction mixture also contains labeled<br />

nucleotide triphosphate.<br />

For assays involving enzymes, a single-stranded unlabeled DNA target is required.<br />

A two-step procedure can be used where the sample is first amplified using the PCR<br />

and the unwanted strand digested away using an exonuclease (11). The primer for the<br />

strand to be retained is synthesized with approximately the last five internucleotide<br />

linkages at the 5′ end as phosphorothioate (12), which protect the strand against<br />

nuclease degradation. After PCR, T7 gene 6 exonuclease is added to the product to<br />

digest away the unwanted, (unphosphorothioated) strand. The resulting single-stranded<br />

DNA can often be used without further purification but, if required, standard purification<br />

techniques could be used, for example, Sephadex or Qiaquick columns.<br />

1.3. Hybridization Assays<br />

Target alleles are distinguished by their ability to hybridize to complementary allele<br />

specific oligonucleotides (ASOs) on the array (Fig. 5A) (8). Hybridization yield and<br />

discrimination depend on temperature, salt concentration, time, and target concentration.<br />

The sequence of the oligonucleotide also affects both yield and discrimination.<br />

Many alleles can be simultaneously examined on the same array, but this requires careful<br />

choice of both oligonucleotide sequence and hybridization and washing conditions<br />

to achieve the maximum discrimination under one set of conditions. Computer-based


Arrays for Genotyping 259<br />

Fig. 3. General scheme for the typing of STRs using oligonucleotide reporter ligation to immobilized oligonucleotides. The addition of<br />

nucleotide triphosphates catalyzed by polymerase can be used in a similar manner.<br />

259


260 Case-Green, Pritchard, and Southern<br />

Fig. 4. Synthesis of an array for typing F13 locus.<br />

methods have proved inadequate for choosing the oligonucleotide sequences because<br />

many factors, such as the position of mismatches and secondary structure in both<br />

the oligonucleotides and target, affect the duplex yield. An empirical solution to<br />

this problem is to test various conditions using combinatorially synthesized arrays<br />

containing many of the possible oligonucleotides complementary to the allele being<br />

investigated (13,14).<br />

1.4. Enzyme-Aided Assays<br />

1.4.1. SNPs<br />

Several variants of the hybridization/extension method are suitable for the analysis<br />

of (biallelic) SNPs (Fig. 5B,C) (15,16). In the simplest form, two oligonucleotides are<br />

synthesized that differ in their terminal base; the two bases complementary to the<br />

bases to be analyzed. Hybridization of the target sequence will occur in similar yields<br />

to both oligonucleotides because of the poor discrimination at terminal bases (17).<br />

Ligases and polymerases can distinguish the two duplexes and are able to extend<br />

only oligonucleotides with the matching base. Labeled nucleotide triphosphate,<br />

complementary to the first base after the allelic position in the target, is incorporated<br />

only on oligonucleotides, which form perfect duplex. The position of the label<br />

on the support indicates the sequence that is complementary to the target (Fig.<br />

5C). Alternatively, a common primer can be used and labeled nucleotide added<br />

complementary to either of the alleles of interest in the target. Only when the nucleotide<br />

added is complementary to the biallelic position will the primer be extended. (Fig.<br />

5B) Ligation can also be used in these assays, replacing the nucleotide triphosphates<br />

with labeled reporter oligonucleotides having terminal bases complementary to the<br />

alleles to be analyzed.<br />

1.4.2. STR Typing<br />

A method for measurement of STR length is shown in Fig. 3. As with SNP typing,<br />

assays using both DNA ligases and DNA polymerases are possible. In the basic ligation<br />

assay, the target sequence plus a reporter oligonucleotide complementary to the distal


Arrays for Genotyping 261<br />

Fig. 5. Three schemes for typing of single nucleotide polymorphisms. The target varies at<br />

the middle of the three bases and can be a G or A. (A) Hybridization to ASOs. (B) Polymerasecatalyzed<br />

extension to a common, locus-specific oligonucleotide primer using nucleotide<br />

triphosphates complementary to one or other of the alleles. In this case, the bases may be<br />

labeled with different fluorophores. (C) Polymerase catalyzed extension to ASO primers using<br />

a common nucleotide triphosphate. The ligation of labeled oligonucleotides catalyzed by ligase<br />

can also be used in a similar manner.


262 Case-Green, Pritchard, and Southern<br />

flanking region to the repeat is hybridized to an array of oligonucleotides that contain a<br />

guide registration sequence plus a varying number of repeats. The ligase can only<br />

join the labeled oligonucleotide to the oligonucleotides of the array where a perfect<br />

duplex junction is formed. This only occurs for oligonucleotides with repeat length<br />

equal to those of the target (18).<br />

The polymerase method is similar. The STRs must have a base of the repeat unit<br />

that differs from the first base in the flanking sequence following the repeats. Labeled<br />

nucleotide triphosphate complementary to the first base after the repeat and DNA<br />

polymerase is added. The labeled nucleotide is incorporated only where the repeat<br />

number on the array matches that of the target. As confirmation, the reaction can be<br />

performed using a labeled nucleotide triphosphate complementary to the first base of<br />

the repeat. Only array oligonucleotides shorter than the target repeat size are extended<br />

in this case.<br />

1.5. Discussion<br />

Three broad classes of assays useful with oligonucleotide arrays are described<br />

above: allele-specific hybridization; primer extension by polymerase (minisequencing);<br />

and ligase assay. All have been used in solution and adapted to arrays. Each assay<br />

incorporates an initial hybridization of target nucleic acid (usually PCR product) to<br />

the oligonucleotide array.<br />

Allele-specific hybridization and the related technique of sequencing by hybridization<br />

for sequencing and resequencing nucleic acids has been under development and in<br />

use for some time (8,14). Hybridization has several well-documented complications, the<br />

major one being the variability of hybridization yield between different oligonucleotide<br />

probes against the nucleic acid target. This variability has two causes, detailed as<br />

follows:<br />

• Sequence-dependent hybridization efficiency. Where oligonucleotides of the array are<br />

of the same length, then the stability of the duplex formed with the nucleic acid target<br />

will depend on the sequence of the oligonucleotides and the type of target (19). This<br />

effect of base composition in DNA/DNA duplexes is lessened by the use of salts, such as<br />

tetramethylammonium chloride as hybridization buffer (20). The stabilities of each set of<br />

duplexes could be measured by thermal denaturation in solution but this is extremely time<br />

consuming and therefore impractical.<br />

• Differences in accessibility of the target to oligonucleotide probes. Secondary and tertiary<br />

structure in both the target and oligonucleotides can prevent hybridization. The problem<br />

can be alleviated by degrading the target to short fragments.<br />

Selecting a set of oligonucleotide probes that give similar hybridization yields<br />

between loci and similar differences in hybridization yield between the allele specific<br />

oligonucleotides at each locus, under the same set of hybridization conditions, is<br />

difficult. In practice, the problem can be minimized by building redundancy into oligonucleotide<br />

arrays (14). Each locus and each allele is represented by oligonucleotides<br />

with different lengths so that at least one pair of allele-specific oligonucleotides will<br />

exhibit the correct hybridization characteristics.<br />

In enzymatic assays, the hybridization step is used to capture the target. This step<br />

can be done at low stringency, allowing efficient hybridization at all ASOs. Alleles are<br />

then discriminated by either polymerase or ligase, which add a labeled reporter group


Arrays for Genotyping 263<br />

to the array at any position where probe/target duplex contains no mismatches. The use<br />

of enzymes that produce a covalently bonded modification to the array also allows the<br />

assays to be thermally cycled. This increases the signal to noise ratio.<br />

Primer extension/minisequencing/genetic bit analysis has the great advantage that the<br />

reporter (labeled dideoxynucleotide triphosphates) is simple and readily commercially<br />

available. At most four different reporters are needed for each assay. This contrasts<br />

with ligase assays where a unique oligonucleotide reporter is required for each locus<br />

studied. Primer extension can also be adapted for resequencing, using a tiling path of<br />

complementary oligonucleotides.<br />

2. Materials<br />

2.1. Array Fabrication<br />

2.1.1. SNP Array Synthesis<br />

1. DNA synthesizer (Perkin–Elmer/ABI).<br />

2. Synthesis cell (Fig. 2, left).<br />

3. Glass or perspex sheet.<br />

4. Deoxynucleotide 3′-DMT, 5′-phosphoramidites (Glen Research).<br />

5. Reagents for oligonucleotide synthesis (Perkin–Elmer/ABI, Pharmacia, Cruachem).<br />

6. Ammonium hydroxide solution (30%) (BDH).<br />

7. Duran bottle.<br />

8. Aminated polypropylene support (Beckman).<br />

9. Scalpel.<br />

2.1.2. STR Array Synthesis<br />

1. See Subheading 2.1.1.<br />

2. Synthesis cell (Fig. 2, right).<br />

2.2. Target Preparation<br />

2.2.1. In Vitro Transcription of RNA<br />

1. PCR reagents.<br />

2. PCR primer for the target containing a T7 or SP6 promoter sequence.<br />

3. T7 or SP6 RNA polymerase (Promega).<br />

4. Transcription buffer 5µ: 200 mM Tris-HCl, pH 7.9, 30 mM MgCl 2 , 10 mM spermidine,<br />

50 mM NaCl; Promega.<br />

5. DTT (100 mM; Promega).<br />

6. RNase Inhibitor (Recombinant RNasin; Promega).<br />

7. [α- 32 P]UTP (3000 Ci/mmol; Amersham).<br />

8. NTPs: ATP, CTP, and GTP as 10 mM stock solutions, and UTP at 250 µmol (all in nuclease<br />

free distilled water; Pharmacia).<br />

2.2.2. Single-Stranded DNA Preparation<br />

1. PCR reagents.<br />

2. PCR primer for the target strand containing 6 phosphorothioate linkages at the 5′-most<br />

6 phosphate linkages.<br />

3. Thermal cycler (MJ Research).<br />

4. T7 gene 6 exonuclease (Amersham).<br />

5. PCR purification kit (optional, Qiagen).


264 Case-Green, Pritchard, and Southern<br />

2.3. Hybridization<br />

2.3.1. Hybridization Assay<br />

1. Buffer: 1 M NaCl, 10 mM Tris-HCl, pH 8.0, 1 mM EDTA, 0.01% SDS.<br />

2. Petri dish.<br />

3. Tissues.<br />

4. Forceps.<br />

5. Strip cut from oligonucleotide array (30 × 2 mm).<br />

6. Target single-stranded RNA (see Subheading 3.2.1.).<br />

2.3.2. Hybridization of Target for Polymerase Assay<br />

1. Buffer: 1 M NaCl.<br />

2. Petri dish.<br />

3. Tissues.<br />

4. Forceps.<br />

5. Strip cut from oligonucleotide array (30 × 2 mm).<br />

6. Target single-stranded DNA (see Subheading 3.2.2.).<br />

2.3.3. Hybridization of Target and Reporter for Ligation Assays<br />

1. Buffer: 1 M NaCl.<br />

2. Petri dish.<br />

3. Tissues.<br />

4. Forceps.<br />

5. Labeled reporter oligonucleotide containing 5′ phosphate group.<br />

6. Target single-stranded DNA (see Subheading 3.2.2.).<br />

7. Strip of oligonucleotide array (30 × 2 mm).<br />

2.4. Enzyme-Aided Assays<br />

2.4.1. Ligation<br />

1. 5µ ligation buffer: 100 mM Tris-HCl, pH 8.3, 0.5% Triton X-100, 50 mM MgCl 2 , 250 mM<br />

KCl, 5 mM NAD + , 50 mM DTT, 5 mM EDTA (Advanced <strong>Bio</strong>technologies Ltd.).<br />

2. Thermus thermophilus DNA ligase (Tth DNA ligase; Advanced <strong>Bio</strong>technologies Ltd).<br />

3. Petri dish set up as humid chamber.<br />

4. Forceps.<br />

2.4.2. Primer Extension with Polymerase<br />

1. 5µ Polymerase buffer: 200 mM Tris-HCl, pH 7.5, 100 mM MgCl 2 , 250 mM NaCl<br />

(Amersham).<br />

2. Polymerase dilution buffer: 50 mM Tris-HCl, pH 7.5; 10 mM DTT (Amersham).<br />

3. DTT (100 mM, Promega).<br />

4. T7 Sequenase V2. 13 U/µL (Amersham).<br />

5. Dideoxynucleotidetriphosphate [α- 33 P]ddNTP (450 µCi/mL; Amersham).<br />

6. Humid chamber.<br />

7. Forceps.<br />

8. Buffer: 10 mM Tris-HCl, pH 8.0, 1 mM EDTA, pH 8.


Arrays for Genotyping 265<br />

3. Methods<br />

3.1. Array Synthesis (Notes 1–3)<br />

3.1.1. SNP Array Synthesis<br />

1. Clamp a reaction cell (Fig. 2, left, 4 × 40 mm) to a piece of aminated polypropylene using<br />

a G-clamp and a glass or perspex backing plate.<br />

2. Attach ports to the synthesizer and synthesize a patch of oligonucleotide corresponding<br />

to the first allele.<br />

3. Displace the cell by 4 mm so that the cell is alongside the previously synthesized patch.<br />

4. Synthesize the oligonucleotide corresponding to the other allele.<br />

5. Repeat as necessary for other alleles.<br />

6. Place the array in a Duran bottle, add 30% ammonia solution, seal the bottle securely,<br />

and heat at 55°C for 6 h.<br />

7. After allowing the bottle and contents to cool, remove the array, wash with ethanol and<br />

dry. Store at –20°C.<br />

8. The array is cut into strips (1- to 2-mm wide) in a direction perpendicular to the stripes of<br />

oligonucleotides using a scalpel and straight edge.<br />

3.1.2. STR Array Synthesis<br />

1. Clamp a cell (Fig. 2, right, 30 × 40 mm) to a piece of aminated polypropylene using a<br />

G-clamp and a glass backing-plate.<br />

2. Attach ports to the synthesizer and synthesize a patch of oligonucleotide corresponding to the<br />

complement of the flanking sequence before the repeats of the target single stranded DNA.<br />

3. On top of this add a sequence complementary to the smallest number of repeats that may<br />

be present in the target DNA.<br />

4. Displace the cell by 2 mm and synthesize a patch corresponding to one repeat.<br />

5. Repeat step 4 until the maximum required number of repeats has been reached.<br />

6. Place the array in a Duran bottle, add 30% ammonia solution, and heat at 55°C for 6 h.<br />

7. After allowing the bottle and contents to cool, remove the array from bottle, wash with<br />

ethanol, and dry. Store at –20°C.<br />

8. The array is cut into strips (1- to 2-mm wide) in a direction perpendicular to the stripes<br />

of oligonucleotides.<br />

3.2. Single-Stranded Target Production<br />

3.2.1. In Vitro Transcription of RNA<br />

1. Perform PCR under conditions that give the required product.<br />

2. Dilute this product to give a concentration of 0.5 to 1.0 µg/µL.<br />

3. In a mircofuge tube at room temperature add the following reagents in order: 5µ transcription<br />

buffer (4 µL); 100 mM DTT (2 µL); RNasin (20 units); 10 mM ATP, CTP, and GTP<br />

(1 µL each); 250 mM UTP (1 µL); Template DNA (2 µL); [α- 32 P] UTP (2 µL) T7 or SP6<br />

RNA polymerase (20 units); and water to give final volume of 20 µL.<br />

4. Mix and Incubate at 37°C for 1 h.<br />

5. Remove unincorporated label by Sephadex G-25 or G-50 chromatography.<br />

3.2.2. Single-Stranded DNA Preparation<br />

1. Perform PCR under conditions that give the required product.<br />

2. Add T7 Gene 6 exonuclease to give a concentration of 2 U per µL.


266 Case-Green, Pritchard, and Southern<br />

3. Incubate at 25°C for 1 h.<br />

4. Incubate at 80°C for 5 min.<br />

5. If desired, purification can be performed using a commercially available purification kit.<br />

3.3. Hybridization (see Note 4)<br />

3.3.1. Hybridization Assay (see Notes 5 and 6)<br />

1. Make up a solution of the target (50–100 fmol, see Subheading 3.2.1.) in 100 µL of<br />

buffer.<br />

2. Soak a tissue in 1X buffer and place around the inner perimeter of a Petri dish to create<br />

a 100% humidity chamber.<br />

3. Place the solution as a puddle in the middle of the Petri dish and place a strip of array<br />

(30 × 2 mm) on top.<br />

4. Incubate for 1hr at the required temperature.<br />

5. Wash the array in fresh buffer solution and blot dry.<br />

3.3.2. Hybridization Reaction for Polymerase Assay (see Note 7)<br />

This method is the same as used in Subheading 3.3.1. except that a solution of<br />

1 to 5 pmol of single-stranded DNA target (see Subheading 3.3.2.) in a 1 M NaCl<br />

solution is made in step 1.<br />

3.3.3. Hybridization for Ligation Assay (see Note 8)<br />

This method is the same as used in Subheading 3.3.1. except that a solution of<br />

1 to 5 pmol of single-stranded DNA target (see Subheading 3.3.2.) and 5 to 10 pmol of<br />

reporter oligonucleotide in a 1 M NaCl solution is made in step 1.<br />

3.4. Enzyme-Catalyzed Assays (see Notes 9–11)<br />

3.4.1. Ligase Assay<br />

1. Make up a solution of DNA ligase (1 U/µL) in 1µ ligase buffer (100 µL) and place as a<br />

puddle in a converted Petri dish.<br />

2. Place a strip of array (30 × 2 mm) that has undergone hybridization according to Subheading<br />

3.3.3. face down in the puddle, ensuring that the surface is completely covered by<br />

the solution.<br />

3. Incubate at 37°C (or 65°C for STR arrays) for 1 to 8 h.<br />

4. Heat in 50/50 water/formamide v/v solution at 100°C for 5 min.<br />

5. Blot dry and image.<br />

3.4.2. Polymerase Assay (see Note 12)<br />

1. Make up 40 µL of a solution containing sequenase (0.5 U/µL), DTT (2.5 mM), polymerase<br />

buffer (1µ), and the appropriate ddNTP (25 nCi/µL).<br />

2. Soak a tissue in water and place round the inside of a Petri dish.<br />

3. Add the solution as a puddle to the middle of the Petri dish.<br />

4. Float a strip of prehybridized array (see Subheading 3.3.2., 1 × 30m) face down on<br />

the solution.<br />

5. Incubate at 37°C for 1 h.<br />

6. Wash away unbound material with Tris-EDTA (TE) buffer at 65°C for 5 min.<br />

7. Blot dry and image.


Arrays for Genotyping 267<br />

4. Notes<br />

1. The synthesis cells are produced by milling or moulding blocks of PTFE. The cavity is<br />

generally 0.5- to 0.75-mm deep and the wall thickness 0.3 to 0.5 mm.<br />

2. Polymerase-catalyzed extension requires the oligonucleotides to be attached through their<br />

5′ ends. For ligation, the oligonucleotides on the array can be attached in either orientation,<br />

using either the method described in Subheadings 2.1. or 2.2. Thus, deoxynucleotide<br />

5′-dimethoxytrityl-3′- phosphoramidites replace the deoxynucleotide 3′-dimethoxytrityl-5′-<br />

phosphoramidites. A requirement for ligation with oligonucleotides is that the 5′ end must be<br />

phosphorylated. This can be achieved by the reaction of a chemical phosphorylating reagent<br />

(Phosphate-On phosphoramidite available from most DNA synthesis reagent suppliers).<br />

This is added as the last step in the chemical synthesis of the oligonucleotides. Phosphate<br />

can also be added post synthetically using polynucleotide kinase and ATP (this reaction<br />

works on both oligonucleotides in solution and array supported oligonucleotides).<br />

3. The use of a linker between the surface and the oligonucleotides improves both hybridization<br />

yield and access to the oligonucleotides by enzymes. Addition of phosphoramidites<br />

based on polyethylene glycol (21) or addition of deoxythymidine phosphoramidites (22)<br />

have both been used to add linkers before oligonucleotide synthesis.<br />

4. Performing hybridizations and reactions of the array strips in puddles allows the volume<br />

of liquid to be minimized. However, the chamber has to be made as humid as possible to<br />

prevent evaporation, especially at elevated temperatures. The methods described use Petri<br />

dishes with wet tissues inside, but the tissues can also be draped over the top of the dish.<br />

The Petri dishes can be placed in sealed plastic containers. Sealed tubes may also be used<br />

and although the volumes are often greater than in puddles of solution, they can be more<br />

easily incubated in hot blocks or in water baths.<br />

Both the hybridization and washing buffer and temperature can be varied. Conditions<br />

should be found that give high yields of duplex and good discrimination between matched<br />

and mismatched duplexes.<br />

5. For hybridizations below 37°C, care must be taken not to warm the plates by touching<br />

with fingers because this can cause the melting of short duplexes. For hybridizations below<br />

room temperature, all the apparatus that comes in contact with the array must be cooled to<br />

the hybridization temperature or below.<br />

6. When using large RNA targets, secondary and tertiary structures can become a problem<br />

preventing hybridization. One method to alleviate this is to randomly cleave the RNA<br />

in base.<br />

7. For assays, including reactions with enzymes, both the buffer and temperature of the<br />

hybridization reaction can be varied. Using a two-step hybridization followed by reaction<br />

procedure the hybridization buffer need not be the buffer required for the enzymatic<br />

reaction. Also, because many of the enzymes used do require specific temperatures, using a<br />

two-step hybridization/reaction the hybridization temperature can be completely different<br />

from the enzyme-catalyzed reaction.<br />

8. The reporter oligonucleotide is typically 5′ end labeled with a radio label using standard<br />

methods. If fluorescence labeling is used, the label is incorporated during oligonucleotide<br />

synthesis.<br />

9. In our methods, radiolabels were used. These are convenient for many applications because<br />

they are readily commercially available and are easily detected using a phosphorimager.<br />

Other label types can be used and in most cases there are similar ready-labeled reagents<br />

available. We have used Cy5 labels on oligonucleotides in ligation assays and nucleotide<br />

triphosphates in polymerase extension assays. This label can be detected using a STORM<br />

phosphorimager. Because of the large fluorescent background of the polypropylene support<br />

at short wavelengths, we have found that fluorescein is not useful with polypropylene.


268 Case-Green, Pritchard, and Southern<br />

10. Instead of performing the assays in two steps, a hybridization followed by a ligation<br />

or polymerization, an all-in-one reaction can be used. There are several advantages to<br />

conducting all-in-one reactions, such as the saving time and effort and the potential to<br />

cycle the temperature allowing the target DNA to be used as a template for more than one<br />

ligation or polymerization. Conducting all-in-one reactions puts additional constraints on<br />

the conditions because the majority of enzymes work in only a small range of temperatures<br />

and salt concentrations. The target single-stranded DNA generally needs to be purified.<br />

11. The methods described use Thermus Thermophilus DNA ligase for ligation reactions<br />

and Sequenase in extension reactions. Other enzyme types can also be used and have<br />

advantages if particular temperatures are required, particularly in all in one reactions.<br />

Therefore, for example, Taq and Thermosequenase have all been used successfully.<br />

12. The extension method described adds a dideoxynucleotidetriphosphate (ddNTP). For most<br />

applications, appropriate deoxynucleotidetriphosphate (dNTP) can be used.<br />

References<br />

1. Case-Green, S. C., Mir, K. U., Pritchard, C. E., and Southern, E. M. (1998) Analysing<br />

genetic <strong>info</strong>rmation with DNA arrays. Curr. Opin. Chem. <strong>Bio</strong>l. 2, 404– 410.<br />

2. Schena, M., Shalon, D., Davis, R. W., and Brown, P. O. (1995) Quantitative monitoring of<br />

gene expression patterns with a complementary DNA microarray. Science 270, 467– 470.<br />

3. Wodicka, L., Dong, H., Mittmann, M., Ho, M. H., and Lockhart, D. J. (1997) Genome-wide<br />

expression monitoring in Saccharomyces cerevisiae. Nat. <strong>Bio</strong>technol. 15, 1359–1367.<br />

4. Milner, N. M., Mir, K. U., and Southern, E. M. (1997) Selecting effective antisense reagents<br />

on combinatorial oligonucleotide arrays. Nat. <strong>Bio</strong>technol. 15, 537–541.<br />

5. Brown, T. and Brown, D. J. S. (1991) Modern machine-aided methods of oligodeoxyribonucleotide<br />

synthesis. In Oligonucleotides and Analogues: A Practical Approach (Eckstein,<br />

R., ed.), IRL Press, Oxford, UK, pp. 1.<br />

6. Southern, E. M., Case-Green, S. C., Elder, J. K., <strong>John</strong>son, M., Mir, K. U., Wang, L., et al.<br />

(1994) Arrays of complementary oligonucleotides for analysing the hybridisation behaviour<br />

of nucleic acids. Nucleic Acids Res. 22, 1368–1373.<br />

7. Matson, R. S., Rampal, J., Pentoney, S. L., Anderson, P. D., and Coassin, P. (1995)<br />

<strong>Bio</strong>polymer synthesis on polypropylene supports: Oligonucleotide arrays. Anal. <strong>Bio</strong>chem.<br />

224, 110–116.<br />

8. Maskos, U. and Southern, E. M. (1993) A novel method for the parallel analysis of multiple<br />

mutations in multiple samples. Nucleic Acids Res. 21, 2269–2270.<br />

9. Matson, R. S., Rampal, B., and Coassin, P. J. (1994) <strong>Bio</strong>polymer synthesis on polypropylene<br />

supports. Anal. <strong>Bio</strong>chem. 217, 306–310.<br />

10. Polymeropoulos, M. H., Rath, D. S., Xiao, H., and Merril, C. R. (1991) Tetranucleotide<br />

repeat polymorphism at the human c-fes/fps proto-oncogene (FES). Nucleic Acids Res.<br />

19, 4018.<br />

11. Nikiforov, T. T., Rendle, R. B., Goelet, P., Rogers, Y.-H., Kotewics, M. L., Anderson,<br />

S., et al. (1994) Genetic bit analysis: a solid phase method for typing single nucleotide<br />

polymorphisms. Nucleic Acids Res. 22, 4167–4175.<br />

12. Zon, G. and Stec, W. J. (1991) In Phosphorothioate oligonucleotides. In Oligonucleotides<br />

and Analogues: A Practical Approach (Eckstein, F., ed.), IRL Press, Oxford, UK, pp. 87.<br />

13. Maskos, U. and Southern, E. M. (1993) A novel method for the analysis of multiple<br />

sequence variants by hybridisation to oligonucleotides. Nucleic Acids Res. 21, 2267–2268.<br />

14. Kozal, M. J., Shah, N., Shen, N., Yang, R., Fucini, R., Merigan, T., et al. (1996) Extensive<br />

polymorphisms observed in HIV-1 clade B protease gene using high-density oligonucleotide<br />

arrays. Nat. Med. 2, 753–759.


Arrays for Genotyping 269<br />

15. Schumaker, J. M., Metspalu, A., and Caskey, C. T. (1996) Mutation detection by solid<br />

phase primer extension. Human Mutat. 7, 346–354.<br />

16. Pastigen, T., Kurg, A., Metspalu, A., Peltonen, L., and Syvanen, A. C. (1997) Minisequencing:<br />

A specific tool for DNA analysis and diagnostics on oligonucleotide arrays. Genome<br />

Res. 7, 606–614.<br />

17. Case-Green, S. C. and Southern, E. M. (1994) Studies on the base pairing properties of<br />

deoxyinosine by solid phase hybridisation to oligonucleotides. Nucleic Acids Res. 22,<br />

131–136.<br />

18. Pritchard, C. E. and Southern, E. M. (1996) Spatially addressable ligation assays: application<br />

of oligonucleotide arrays to DNA fingerprinting. In Innovation and Perspectives in<br />

Solid Phase Synthesis and Combinatorial Libraries (Epton, R., ed.), Mayflower Scientific,<br />

Birmingham, UK, pp. 499.<br />

19. Ratmeyer, L., Vinayak, R., Zhong, Y. Y., Zon, G., and Wilson, W. D. (1994) Sequence<br />

specific thermodynamic and structural properties for DNA-RNA duplexes. <strong>Bio</strong>chemistry<br />

33, 5298–5304.<br />

20. Maskos, U. and Southern, E. M. (1993) A study of oligonucleotide reassociation using<br />

large arrays of oligonucleotides synthesised on a glass support. Nucleic Acids Res. 21,<br />

4663– 4669.<br />

21. Shchepinov, M. S., Case-Green, S. C., and Southern, E. M. (1997) Steric factors influencing<br />

hybridisation of nucleic acids to oligonucleotide arrays. Nucleic Acids Res. 25,<br />

1151–1161.<br />

22. Guo, Z., Guilfoyle, R. A., Thiel, A. J., Wang, R., and Smith, L. M. (1994) Direct fluorescence<br />

analysis of genetic polymorphisms by hybridization with oligonucleotide arrays on glass<br />

supports. Nucleic Acids Res. 22, 5456–5465.


270 Case-Green, Pritchard, and Southern


Serial Analysis of Gene Expression 271<br />

40<br />

Serial Analysis of Gene Expression<br />

Karin A. Oien<br />

1. Introduction<br />

Serial analysis of gene expression (SAGE ) is a patented, large-scale mRNAprofiling<br />

technology that produces comprehensive, quantitative, and reproducible<br />

gene expression profiles (originally described in refs. 1 and 2). Unlike the alternative<br />

technologies of differential display and subtractive hybridization, SAGE produces a<br />

full catalog of all transcripts, not only differentially expressed genes, and unlike smaller<br />

arrays, SAGE needs no assumptions about the genes that are likely to be expressed,<br />

thus allowing the identification of novel genes (are among many excellent reviews, see<br />

refs. 3–5). SAGE is based on generating clones of concatenated (linked) short sequence<br />

tags derived from mRNA from the target cells or tissue (Fig. 1).<br />

Each tag is 9- or 10-bp long and represents one mRNA and each clone insert contains<br />

up to 40 tags joined serially. Therefore, sequencing of multiple concatenates describes<br />

the pattern and abundance of mRNA, with an improvement in efficiency of up to<br />

40-fold compared with conventional analysis of expressed sequence tags. The mRNA<br />

transcript corresponding to the short SAGE tag is identified from genetic databases<br />

using appropriate software.<br />

The SAGE method was developed in 1995 by Velculescu et al. (1) of the Kinzler and<br />

Vogelstein laboratory at <strong>John</strong>s Hopkins University, from where the SAGE protocol,<br />

software, and updates are available to academic investigators for noncommercial use<br />

via http://www.sagenet.org/sage_protocol.htm. For commercial purposes, the user<br />

should contact Genzyme Corporation, who own and license the SAGE technology.<br />

This and other useful web sites are listed in Table 1. Since the first report of SAGE,<br />

many technical modifications have been described. Some enable the use of smaller<br />

amounts of starting material (3,6–10), whereas others have improved the efficiency<br />

of intermediary SAGE reactions (11–14). Invitrogen has also released a kit called<br />

I-SAGE , which provides all of the numerous reagents required in a high-quality<br />

form.<br />

The <strong>John</strong>s Hopkins protocol is 27 pages long, and Invitrogen’s online I-SAGE <br />

instructions contain 73 pages. This is because SAGE involves many stages, each using<br />

well-established and relatively straightforward molecular biological techniques, but any<br />

of which can go wrong! This article thus provides concise instructions with an emphasis<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

271


272 Oien<br />

Fig. 1. A schematic diagram of the SAGE method: the numbers refer to the corresponding<br />

stages in Subheading 3.<br />

on common problems (well discussed in ref. 15), but it is strongly recommended<br />

that one also consult the full protocols and references supplied in this chapter.<br />

The software used to analyze the tags has been regularly updated, and alternative<br />

programs have been developed, including eSAGE (16), USAGE (17), and ExProView<br />

(18). The web-based bio<strong>info</strong>rmatics facilities of the National Center for <strong>Bio</strong>technology<br />

Information (NCBI) are extremely useful (19). There, through SAGEmap (20),<br />

numerous publicly available SAGE libraries can be accessed online both for global


Serial Analysis of Gene Expression 273<br />

Table 1<br />

Useful Web Sites for SAGE<br />

URL<br />

Contents<br />

http://www.sagenet.org/sage_protocol.htm <strong>John</strong>s Hopkins SAGE protocol and software<br />

for non-commercial use, plus conferences,<br />

publications and other <strong>info</strong>rmation (1,2)<br />

http://www.genzyme.com/sage/<br />

Genzyme Molecular Oncology: <strong>info</strong>rmation<br />

on SAGE and its commercial use<br />

http://www.invitrogen.com<br />

Invitrogen including I-SAGE protocol<br />

http://www.dsv.cea.fr/thema/get/<br />

SADE: a SAGE Adaptation for Downsized<br />

sade.html Extracts (6)<br />

http://www.ambion.com<br />

Ambion website with advice on RNA work<br />

http://www.ncbi.nlm.nih.gov/SAGE/ NCBI’s SAGEmap (20)<br />

http://www.ncbi.nlm.nih.gov/UniGene/<br />

NCBI’s UniGene (“unique gene”: clusters all<br />

index.html transcripts of one gene under one name) (19)<br />

http://www.geneontology.org<br />

Gene Ontology databases: <strong>info</strong>rmation on<br />

gene function and cellular location (25)<br />

comparisons and for investigation of the expression of individual transcripts, which can<br />

be further studied via Unigene (see Table 1, ref. 19). The statistical basis for designing<br />

and analyzing SAGE experiments has also been discussed in detail (21–24). Here,<br />

however, the focus is on the wet laboratory work.<br />

2. Materials<br />

The reagents specified have been used successfully, but many equivalents are<br />

available and could be used if preferred (see Note 1). This description is based on<br />

the <strong>John</strong>s Hopkins protocol (version 1.0d); the most helpful modifications are also<br />

described.<br />

2.1. RNA Work (see Note 2)<br />

1. Aerosol-resistant pipet tips (Greiner Labortechnik Ltd, Gloucestershire, UK).<br />

2. DEPC-treated water (Diethylpyrocarbonate: Sigma, Dorset, UK).<br />

3. (Optional) RnaseZap ® (Ambion (Europe) Ltd, Cambridgeshire, UK).<br />

2.2. Kits for Purification of RNA and cDNA Synthesis<br />

1. Preparation of total RNA: TRIZOL ® Reagent (Invitrogen Life Technologies, Paisley, UK).<br />

2. Purification of polyA+ mRNA: Poly(A)Purist Kit (Ambion).<br />

3. (Alternative protocols: streptavidin-coated tubes obtained with mRNA Capture Kit or<br />

separately (Roche Diagnostics Ltd, East Sussex, UK), instead of steps 1 and 2 (see Note 3<br />

and Subheadings 2.4. and 2.6.).<br />

4. cDNA synthesis: SUPERSCRIPT Choice System for cDNA Synthesis (Invitrogen).<br />

2.3. Purification of DNA<br />

1. Phenolchloroform (P/C): phenolchloroformisoamyl alcohol (25241; Ambion).<br />

2. Glycogen, molecular biology grade (20 mg/mL) (Roche Diagnostics Ltd).<br />

3. Ammonium acetate (10 M).


274 Oien<br />

4. 100% and 70% ethanol.<br />

5. LoTE: 3 mM Tris-HCl (pH 7.5), 0.2 mM EDTA (pH 7.5).<br />

6. TE: 10 mM Tris-HCl (pH 7.5), 1 mM EDTA (pH 7.5).<br />

7. (Optional) QIAquick ® DNA Cleanup System: QIAquick Nucleotide Removal and PCR<br />

Purification Kits; and, for higher through-put, QIAquick 8 PCR Purification Kit with<br />

QIAvac 6S (Qiagen Ltd, West Sussex, UK).<br />

2.4. Oligonucleotides<br />

1. <strong>Bio</strong>tinylated oligo dT: 5′ [biotin] T 18 (Alternative protocols, see Note 3).<br />

2. SAGE Linker 1A: 5′ TTT GGA TTT GCT GGT GCA GTA CAA CTA GGC TTA ATA<br />

GGG ACA TG 3′<br />

3. SAGE Linker 1B: 5′ TCC CTA TTA AGC CTA GTT GTA CTG CAC CAG CAA ATC C<br />

[amino mod. C7] 3′<br />

4. SAGE Linker 2A: 5′ TTT CTG CTC GAA TTC AAG CTT CTA ACG ATG TAC GGG<br />

GAC ATG 3′<br />

5. SAGE Linker 2B: 5′ TCC CCG TAC ATC GTT AGA AGC TTG AAT TCG AGC AG<br />

[amino mod. C7] 3′<br />

6. SAGE Primer 1: 5′ [biotin] GGA TTT GCT GGT GCA GTA CA 3′ (11)<br />

7. SAGE Primer 2: 5′ [biotin] CTG CTC GAA TTC AAG CTT CT 3′ (11)<br />

8. M13 Forward: 5′ GTA AAA CGA CGG CCA GT 3′<br />

9. M13 Reverse: 5′ GGA AAC AGC TAT GAC CAT G 3′<br />

The working concentration of all primers, except biotinylated oligo dT, is 350 ng/µL.<br />

Obtain linkers and biotinylated oligos gel-purified: recommended suppliers include<br />

Integrated DNA Technologies, Inc. (IA, USA) and Oswel Research Products Ltd.<br />

(Southampton, UK). Before use, linkers must be kinased, either biochemically during<br />

synthesis or later enzymatically (see Note 4), and annealed to the complementary<br />

linker.<br />

2.5. Restriction and Modifying Enzymes<br />

1. NlaIII (10 U/µL), supplied with NEBuffer 4 and 100× bovine serum albumin (New<br />

England <strong>Bio</strong>labs [NEB] Inc, MA,) stored at –70°C).<br />

2. BsmFI (2 U/µL) (NEB).<br />

3. SphI (5 U/µL) (NEB).<br />

4. T4 Polynucleotide Kinase 10U/µL (NEB) (or obtain linker oligonucleotides already<br />

kinased, see Note 4).<br />

5. T4 DNA Ligase (5 U/µL, or 20 U/µL if preferred; NEB). Aliquot ligase enzyme and buffer<br />

to prevent cross-contamination from PCR products.<br />

6. DNA polymerase I large fragment (Klenow); 5 U/µL (NEB).<br />

2.6. Magnetic Beads and Related Materials<br />

1. Dynabeads ® M-280 Streptavidin (-coated) (Dynal AS, Oslo, Norway).<br />

2. Magnetic particle concentrator (magnetic stand to immobilize beads) and sample mixer<br />

(Dynal AS).<br />

3. 2× Binding and washing (B+W) buffer: 2 M NaCl, 10 mM Tris-HCl (pH 7.5), 1 mM<br />

EDTA. Also prepare 1× B+W buffer solution.<br />

4. (Optional) Nonstick RNase-free tubes (Ambion) have been recommended for use with<br />

magnetic beads to minimize adhesion of beads to the tube walls.<br />

5. (Alternative protocols: Dynabeads ® Oligo dT 25 (Dynal AS; see Note 3 and Subheadings<br />

2.2. and 2.4.).


Serial Analysis of Gene Expression 275<br />

2.7. Gel Materials<br />

1. 40% polyacrylamide (37.51 acrylamidebis) (<strong>Bio</strong>–Rad Ltd, Hertfordshire, UK).<br />

2. 40% polyacrylamide (191 acrylamidebis) (<strong>Bio</strong>–Rad).<br />

3. DNA molecular weight markers, for example, 10- and 100-bp DNA ladders (Invitrogen).<br />

4. SYBR ® Green I nucleic acid gel stain (Molecular Probes Inc., OR) or a 0.5 µg/mL solution<br />

of Ethidium Bromide (Sigma).<br />

5. Gel-loading buffer containing Bromophenol Blue and Xylene Cyanol (e.g., Ambion).<br />

6. Apparatus for vertical polyacrylamide gel electrophoresis, for example, Atto Maxi Slab<br />

with 16- × 16-cm glass gel plates, 1.5-mm spacers, and 12- and 20-well combs (Genetic<br />

Research Instrumentation, Essex, UK).<br />

7. 50× Tris-Acetate buffer: 242 g of Tris base, 57.1 mL of glacial acetic acid, and 100 mL of<br />

0.5 M EDTA (pH 8.0) made up to 1 L with dH 2 O.<br />

8. 10% Ammonium persulfate.<br />

9. TEMED (N,N,N′,N′-Tetramethylethylenediamine; Sigma).<br />

10. Sterile scalpel.<br />

11. 21-gauge needles.<br />

12. Spin-X filter microcentrifuge tubes (Corning Ltd, Buckinghamshire, UK).<br />

13. Materials, buffers, and apparatus for agarose gel electrophoresis<br />

2.8. Polymerase Chain Reaction (PCR)<br />

1. Option 1: <strong>John</strong>s Hopkins protocol. PLATINUM Taq DNA Polymerase (5 U/µL; Invitrogen);<br />

10× PCR buffer: 166 mM (NH 4 ) 2 SO 4 , 670 mM Tris-HCl (pH 8.8), 67 mM MgCl 2 , 100 mM<br />

β-mercaptoethanol; Dimethyl sulphoxide (Sigma). Option 2: own modification. HotStarTaq<br />

DNA Polymerase Kit and Taq PCR core kit (Qiagen).<br />

2. dNTPs (10 mM; Invitrogen; included in some kits).<br />

3. Thermal cycler of choice (e.g., Thermo Hybaid, Middlesex, UK).<br />

4. Mineral oil (Sigma). Ensure that it is free from PCR contamination.<br />

2.9. Cloning<br />

1. For cloning of concatemers: Zero Background Cloning Kit (Invitrogen).<br />

2. ELECTROMAX DH10B Cells (Invitrogen).<br />

3. S.O.C. Medium (Invitrogen).<br />

4. Low-salt LB medium and agar plates with Zeocin , prepared according to instructions<br />

in Zero Background Cloning Kit.<br />

5. Electroporator and electroporation cuvettes (Thermo Hybaid).<br />

2.10. Sequencing<br />

1. Sequencing as performed locally, for example, use BigDye Primer Kit and ABI automated<br />

sequencer (both Perkin–Elmer from Applied <strong>Bio</strong>systems, Warrington, UK).<br />

2.11. SAGE <strong>Bio</strong><strong>info</strong>rmatics<br />

<strong>John</strong>s Hopkins SAGE program (downloaded from web site after registration), in<br />

combination with NCBI’s GenBank ® databases and Microsoft ® Access and Excel<br />

programs, or other software (see Subheading 1., last paragraph, and Table 1).<br />

3. Methods (see Notes 1 and 2)<br />

3.1. mRNA Preparation<br />

1. Prepare total RNA from cells or tissue by standard methods, for example, TRIZOL<br />

Reagent.


276 Oien<br />

2. Purify polyA + mRNA from total RNA using, for example, Poly(A)Purist kit. This original<br />

protocol requires 2.5 to 5 µg of mRNA for the next step (but see Note 3).<br />

3.2. cDNA Synthesis<br />

1. Synthesize double-stranded cDNA using SUPERSCRIPT ® Choice System. For the first strand<br />

synthesis, use 2.5 µg of the biotinylated oligo dT purchased separately, instead of that<br />

supplied with the kit. Synthesize the second strand, proceed to the next step in this SAGE<br />

protocol. (Do not perform EcoRI adapter addition).<br />

2. Phenolchloroform (P/C) extract and ethanol precipitate (see Notes 5–7). Resuspend<br />

in 20 µL of LoTE.<br />

3.3. Cleavage of <strong>Bio</strong>tinylated cDNA with Anchoring Enzyme NlaIII<br />

to Create CATG Sticky-End<br />

1. To 10 µL (half) of the biotinylated cDNA, add 74 µL of LoTE, 10 µL of 10× NEBuffer 4,<br />

1 µL of 100× BSA, and 5 µL of NlaIII. Incubate at 37°C for 1 h.<br />

2. P/C extract, ethanol precipitate, and resuspend in 20 µl LoTE.<br />

3.4. Binding of <strong>Bio</strong>tinylated cDNA to Magnetic Beads<br />

1. Add 100 µL of streptavidin Dynabead slurry to each of two 1.5-mL microcentrifuge tubes.<br />

Immobilize beads with Dynal magnet and remove supernatant.<br />

2. Wash beads with 200 µL of 1× B+W buffer then remove supernatant.<br />

3. Add 100 µL of 2× B+W buffer, 90 µL of dH 2 O, and 10 µL of cleaved biotinylated cDNA to<br />

each of the two tubes. Incubate with gentle mixing for 15 min at room temp.<br />

4. Immobilize beads with magnet and remove supernatant. Wash three times with 200 µL of<br />

1× B+W buffer and once with 200 µL of LoTE. Proceed immediately to next step.<br />

3.5. Ligating CATG Sticky-Ended Linkers to Bound cDNA<br />

1. Linkers must be kinased and annealed in advance (see Note 4).<br />

2. Use one of the two tubes containing washed magnetic beads for linker 1 and the other for<br />

linker 2. Immobilize beads with magnet and remove supernatant.<br />

3. Add 29 µL of LoTE, 5 µL of annealed linker 1 or 2, and 4 µL of 10× ligase buffer.<br />

Resuspend the beads with gentle mixing (flick tube with finger).<br />

4. Heat at 50°C for 2 min then allow to cool to room temp over 15 min.<br />

5. Add 2 µL of T4 DNA Ligase. Incubate at 16°C for 2 h with intermittent gentle mixing.<br />

6. After ligation, wash beads three times with 200 µL of 1× B+W buffer. Transfer to new tubes.<br />

Wash once with 200 µL of 1× B+W buffer and twice with 200 µL of 1× NEBuffer 4.<br />

3.6. Creation of cDNA Tags and Their Release<br />

from Magnetic Beads Using Tagging Enzyme BsmFI<br />

1. Remove buffer then add 87 µL of LoTE, 10 µL of 10× NEBuffer 4, and 1 µL of 100× BSA<br />

to each tube. Pre-incubate at 65°C for 2 min.<br />

2. Add 2 µL of BsmFI. Incubate at 65°C for 1 h with intermittent gentle mixing.<br />

3. Immobilize beads with magnet. This time, collect supernatant and transfer to two new<br />

tubes. Wash beads with 100 µL of 1× NEBuffer 4. Collect buffer and add to previous<br />

supernatant, discarding beads.<br />

4. P/C extract, ethanol precipitate, and resuspend in 10 µL of LoTE.


Serial Analysis of Gene Expression 277<br />

3.7. Blunt-Ending-Released cDNA Tags<br />

1. To each of the two tubes (from linkers 1 and 2) containing released cDNA tags, add 31 µL<br />

dH 2 O, 5 µL of 10× EcoPol buffer, 0.5 µL of 100× BSA, 2.5 µL of 10 mM dNTPs, and<br />

1 µL of DNA polymerase I large fragment (Klenow).<br />

2. Incubate at 37°C for 30 min, and then pool both tubes of blunt-ended tags.<br />

3. P/C extract, ethanol precipitate, and resuspend in 12 µL of LoTE.<br />

3.8. Ligating Blunt-Ended Tags to Form 102-bp Ditags<br />

1. Set up two new tubes, ideally of small size (0.2 mL). One tube is for the ditag ligation<br />

reaction. The other is a negative ligation control, to exclude cross-contamination at the<br />

next, PCR, step: set this tube up first. To each of the two tubes, add 4 µL of blunt-ended<br />

tags, 0.8 µL dH 2 O and 0.6 µL of 10× ligase buffer.<br />

2. To the negative control tube, add 0.6 µL of dH 2 O.<br />

3. To the ditag reaction, add 0.6 µL of T4 DNA ligase.<br />

4. Cover with a drop of mineral oil to avoid evaporation and incubate at 16°C overnight.<br />

3.9. PCR Amplification of 102-bp Ditags<br />

The PCR aims to produce sufficient 102-bp ditag DNA for subsequent isolation and<br />

concatemerization of 26 bp ditags but may itself be problematic (see Note 8).<br />

1. After ligation, add 14 µL of LoTE to increase volume to 20 µL and mix. Take 1 µL<br />

and dilute 100-fold with LoTE. Use 1 µL of the dilution in a 50- or 100-µL PCR with<br />

biotinylated SAGE Primers 1 and 2 (11). To avoid cross-contamination, set up the two<br />

negative control reactions (no template and no ligase) first.<br />

2. Step 2, option 1: <strong>John</strong>s Hopkins protocol. Per 50-µL reaction, use 30.5 µL dH 2 O, 5 µL<br />

of 10× SAGE PCR buffer, 3 µL of DMSO, 7.5 µL of 10mM dNTPs, 1 µL of each of<br />

SAGE Primers 1 and 2, and 1 µL PLATINUM Taq DNA Polymerase. The cycling parameters,<br />

optimized for a Hybaid thermal cycler, are as follows: 94°C for 1 min; 26 to 30 cycles of<br />

94°C for 30 s, 55°C for 1 min and 70°C for 1 min; then 70°C for 5 min. Optimize with<br />

different template dilutions (1/50, 1/100, or 1/200 per reaction). Step 2, option 2: own<br />

modification. The HotStarTaq DNA Polymerase Kit routinely works well with 1 µL, or<br />

often less, of the 1/100 dilution with no need for further adjustment of template dilution.<br />

Per 100-µL reaction, use: 60.5 µL of dH 2 O, 10 µL of 10× Qiagen PCR buffer, 5 µL of<br />

25 mM MgCl 2 , 20 µL of 5× Q-Solution, 2 µL of 10 mM dNTPs, 0.5 µL of each of SAGE<br />

Primers 1 and 2, and 0.5 µLHotStarTaq DNA Polymerase. The cycling parameters are:<br />

94.5°C for 15 min; 26 to 30 cycles of 94.5°C for 30 s; 56°C for 1 min and 72°C for 1 min;<br />

then 72°C for 5 min.<br />

3. Optimize the cycle numbers between 26 and 30. More than 30 cycles usually results in<br />

high molecular weight smearing with less of the desired product.<br />

4. After PCR, load 10 µL of each reaction on a 12% polyacrylamide gel with a 20-bp ladder.<br />

Run at 160 V for 2.5 h until the bromophenol blue dye front has run out of the gel and<br />

the xylene cyanol is 1 to 2 cm from the bottom, then stain (see Note 9). The amplified<br />

ditags should produce a 102-bp band. Background bands are common: the brightest runs<br />

at 80 bp and contains amplified ligated linkers without tags. The negative controls should<br />

contain no product.<br />

5. After optimization, perform large-scale PCR by preparing then distributing a master-mix<br />

into three 96-well PCR plates with 50 or 100 µL in each well.<br />

6. After PCR, pool the reactions. P/C extract and ethanol precipitate, scaling up as needed.<br />

The large volumes can be dealt with using either multiple 1.5-mL microcentrifuge tubes or


278 Oien<br />

50-mL conical tubes. Resuspend in a total of 250 µL of LoTE. (see Note 10 for optional<br />

DNA quantitation at this stage.)<br />

3.10. Isolation of 102-bp Ditags by Gel Purification<br />

1. Load pooled PCR products on three 12% polyacrylamide gels. Run and stain as before.<br />

2. Cut out the 102-bp band of amplified ditags.<br />

3. Fragment gel by placing cut-out bands in 0.5 mL of microcentrifuge tubes that have previously<br />

been pierced through the base with a needle, and insert into a 2-mL tube. (Depending<br />

on bandwidth, 3–4 tubes are used per gel.) Centrifuge at full speed for 2 min.<br />

4. Elute DNA from gel fragments by adding 250 µL of LoTE and 50 µL of 10 M ammonium<br />

acetate to each 2-ml tube. Ensure that gel fragments are covered by buffer (add more if<br />

necessary). Vortex tubes then incubate at 65°C for 2 h.<br />

5. For each 0.5-mL tube, prepare two Spin-X filter microcentrifuge tubes by placing 5 µL of<br />

LoTE on the filter. Transfer contents of each 0.5-mL tube to two Spin-X tubes. Centrifuge<br />

at full speed for 5 min. (see Note 7 and ref. 12).<br />

6. Pool samples and ethanol precipitate. Resuspend in a total of 100 µL of LoTE.<br />

3.11. Isolation of 26-bp Ditags by NlaIII Digestion and Removal<br />

of Linkers Using Magnetic Beads and Gel Purification<br />

1. To the pooled PCR products, add 58 µL of LoTE, 20 µL of 10× NEBuffer 4, 2 µL of 100×<br />

BSA and 20 µL of NlaIII. Incubate at 37°C for 1 h.<br />

2. During NlaIII digestion, add 100 µL of streptavidin Dynabeads to each of two 1.5-mL<br />

tubes. Immobilize beads with magnet and remove supernatant. Wash beads with 200 µL of<br />

1× B+W buffer. Remove buffer when NlaIII digestion is complete.<br />

3. To each of the two tubes containing streptavidin Dynabeads, add 100 µL of 2× B+W buffer<br />

and 100 µL (half) of the NlaIII digestion (11).<br />

4. Incubate with gentle mixing for 15 min at room temp.<br />

5. Immobilize beads with a magnet. Collect supernatant.<br />

6. Wash beads once with 200 µL of 1× B+W buffer and once with 200 µL of LoTE. Collect<br />

supernatant in each case.<br />

7. Pool supernatants and keep on ice to prevent ditag denaturation (13). Discard beads.<br />

8. P/C extract at 4°C. Ethanol precipitate with centrifugation at 4°C (13). Resuspend in 15 µL<br />

of TE (not LoTE).<br />

9. Load on two lanes of a 12% polyacrylamide gel with a 20-bp ladder. Run at 160 V for 2 h<br />

until the bromophenol blue dye front is 3 cm from the bottom. Stain. (Purified ditags<br />

run at 22–26 bp. Released linkers run at 40 bp. Bands between 60–100 bp result from<br />

incomplete NlaIII digestion.)<br />

10. Cut out ditag band running at 22–26 bp.<br />

11. Elute DNA from gel fragments as before, but incubate at 37°C, not 65°C.<br />

12. Ethanol precipitate with centrifugation at 4°C (13). Resuspend in 6.4 µL of LoTE.<br />

3.12. Ligation of Sticky-Ended 26-bp Ditags to Form Concatemers<br />

Then Gel Purification of Concatemers<br />

1. To the purified ditags, add 0.8 µL of 10× ligase buffer and 0.8 µL of T4 DNA Ligase.<br />

Incubate at 16°C for 2 h (or longer, e.g., overnight, if desired).<br />

2. Add loading buffer directly to concatemer ligation. Heat at 65°C for 15 min then chill on<br />

ice for 5 min (14). Load on 8% polyacrylamide gel in one lane with a 100-bp ladder. Run at<br />

130 V for 3 h, until the bromophenol blue is 3 cm from the bottom. Stain.<br />

3. Cut out DNA smear over 500 bp in size.<br />

4. Elute concatemer DNA from gel fragments as before but incubate at 65°C.<br />

5. Ethanol precipitate and resuspend in 6 µL of LoTE.


Serial Analysis of Gene Expression 279<br />

3.13. Cloning Concatemers<br />

Clone concatemers using the Zero Background Cloning Kit. The pZErO ® -1 vector<br />

contains a lethal gene which is disrupted by insertion of DNA, thus only positive<br />

recombinants should grow (this is the theory: in practice, some colonies do lack inserts).<br />

1. To linearize the vector, mix 1 µL of pZErO ® -1 (1 µg/µL), 7 µL of dH 2 O, 1 µL of NEBuffer<br />

2 and 1 µL of SphI. Incubate at 37°C for 15 to 30 min (not over 30 min).<br />

2. P/C extract and ethanol precipitate. Resuspend in 30 µL of LoTE.<br />

3. To 1 µL of SphI-linearized pZErO ® , add the 6 µL of purified concatemers, 1 µL of<br />

10× ligase buffer and 1 µL of T4 DNA Ligase. Include no insert (omit concatemers) and no<br />

ligase (omit concatemers and ligase) controls. Incubate at 16°C for 2 h.<br />

4. P/C extract and ethanol precipitate. Resuspend in 6 µL of LoTE.<br />

5. Transfect 2 µL of DNA into ELECTROMAX DH10B Escherichia coli cells by electroporation,<br />

according to manufacturer’s instructions.<br />

6. Plate one-tenth of transfected bacteria onto each 13-cm Zeocin -containing low-salt LB<br />

agar plate. Keep all plates at 4°C until inserts are checked. If inserts are of appropriate<br />

size, plates may be used for large-scale sequencing.<br />

3.14. Screening of Transformants by PCR<br />

to Identify Long Concatemer Inserts<br />

Perform PCR with vector-specific primers to determine insert size in each bacterial<br />

colony. Tubes (0.5 mL) may be used initially, but 96-well PCR plates are more efficient<br />

for large-scale screening. The 25-µL reaction volume can be reduced, for example,<br />

to 16 µL.<br />

1. Step 1, option 1: <strong>John</strong>s Hopkins protocol. Per 25 µL of reaction, use: 2.5 µL of 10× SAGE<br />

PCR buffer; 1.25 µL of DMSO; 1.25 µL of 10 mM dNTPs; 0.5 µL of each of M13 forward<br />

and reverse Primers; 19 µL of dH 2 O; and 0.2 µL of PLATINUM Taq DNA Polymerase. The<br />

cycling parameters are 95°C for 2 min; 25 cycles of 95°C for 30 s, 56°C for 1 min and<br />

70°C for 30 s; then 70°C for 5 min. Step 1, option 2: own modification. The colony PCRs<br />

are robust and work well with the Taq PCR core kit. Per 100 µL of reaction volume, use:<br />

61.5 µL of dH 2 O, 10 µL of 10× Qiagen PCR buffer, 5 µL of 25 mM MgCl 2 , 20 µL of<br />

5× Q-Solution, 2 µL of 10 mM dNTPs, 0.5 µL of each of M13 forward and reverse Primers,<br />

and 0.5 µL of Taq DNA Polymerase. The cycling parameters are 94.5°C for 1.5 min;<br />

30 cycles of 94.5°C for 30 s, 52°C for 1 min and 72°C for 1 min; then 72°C for 5 min.<br />

2. Touch a single colony with a new pipet tip then dip into reaction mix and shake. Repeat<br />

as necessary. Perform PCR.<br />

3. Run 5 µL of each reaction on a 2% agarose gel with a 100-bp ladder.<br />

3.15. Sequencing of SAGE Concatemer Inserts<br />

Select and sequence the PCR products of 500 bp in size or over because these<br />

should contain at least 15 tags (226 bp of flanking pZErO ® -1 vector plus 12 to 13 bp<br />

per tag).<br />

1. Before sequencing, purify the PCR product (partly to remove primers because M13<br />

forward is used again). Individual phenol chloroform extraction and ethanol precipitation<br />

is one option. However, Qiagen’s QIAquick 8 PCR Purification Kit with the QIAvac 6S<br />

are more efficient for large-scale sequencing (see Note 7).<br />

2. Sequence according to local preference. For example, use the BigDye Primer Kit with<br />

one-half to one-tenth of the purified PCR product per sequencing reaction and the M13


280 Oien<br />

Forward Primer, and then run the reaction on an ABI automated sequencer. In some<br />

laboratories, DNA purification and sequencing can be fully automated.<br />

3.16. Analysis of SAGE Sequence Files<br />

Within the concatemer sequences, the linked ditags of approx 26 bp are separated<br />

by CATG, which is the recognition site of NlaIII. The SAGE software uses the CATG<br />

sequence to identify and extract the ditags, which are then halved into individual<br />

tags. The software then quantifies the number of times the tag occurs within a given<br />

population of clone inserts and creates a report of the abundance of each tag. The report<br />

can be linked to genetic databases for identification of the gene(s) corresponding to the<br />

tags and used to compare different SAGE libraries.<br />

Use the downloaded <strong>John</strong>s Hopkins SAGE software according to the instructions<br />

provided, in combination with NCBI’s Genbank databases, and Microsoft Access and<br />

Excel. Other SAGE programs are also available (see Subheading 1 and Table 1).<br />

3.17. Discussion<br />

The I-SAGE instructions suggest that the whole procedure takes at least 9 d and<br />

longer to screen and sequence the selected clones. A few weeks, or more likely, months,<br />

is a more realistic estimate, especially if setting up SAGE on your own without a<br />

kit. Here I have emphasized commonly encountered problems, but the more detailed<br />

<strong>John</strong>s Hopkins and I-SAGE protocols contain excellent trouble-shooting sections, and<br />

I-SAGE in particular describes verification steps to check the success of each stage.<br />

Once the SAGE libraries have been produced and analyzed, individual SAGE tags<br />

may be selected for further study, through NCBI’s SAGEmap (20) and Unigene (19),<br />

and the developing Gene Ontology databases (25), among other resources. This process<br />

is straightforward where the tag clearly corresponds to one gene but may be more<br />

difficult where either no matching gene or multiple matches exist. This problem can be<br />

addressed by reverse-transcription PCR with the short SAGE tag as a primer (26–28).<br />

This generates longer, more specific, 3′ cDNA fragments that facilitate investigation of<br />

the gene and can also be used to check whether the tag is truly differentially expressed<br />

between samples of interest (26).<br />

To conclude, SAGE is an excellent method of mRNA expression profiling. Although<br />

it is time-consuming and laborious, and requires expertise in molecular biology, the<br />

resulting libraries are extremely valuable, providing data that are truly comprehensive<br />

and quantitative and that enable the identification of novel genes.<br />

4. Notes<br />

1. General instructions. These concise instructions assume knowledge of and experience<br />

in standard molecular biological techniques: purification and manipulation of RNA and<br />

DNA, including phenol:chloroform extraction and ethanol precipitation; cDNA synthesis;<br />

restriction enzyme digestion; PCR; agarose and polyacrylamide gel electrophoresis;<br />

cloning; and sequencing. For further <strong>info</strong>rmation, consult the references and web sites<br />

listed and, of course, Sambrook et al. (29). Where kits are used, follow the manufacturer’s<br />

protocol. Most other reagents, including enzymes and magnetic beads, are also supplied<br />

with detailed instructions. Alternatively, use Invitrogen’s I-SAGE kit and protocol<br />

throughout. The availability of standard laboratory equipment is also assumed: 0.2-, 0.5-,<br />

1.5-, and 2- mL microcentrifuge tubes; 96-well PCR plates; a microcentrifuge, preferably


Serial Analysis of Gene Expression 281<br />

refrigerated; a vortex mixer; 50-mL tubes and an appropriate centrifuge; wet and dry ice;<br />

water baths; and incubators, including one for bacterial culture.<br />

2. Working with RNA. The quality and quantity of input RNA is critical to the success or<br />

failure of SAGE. Working with RNA may be difficult, and advice is given in Maniatis<br />

and on Ambion’s web site. Materials should be RNase-free: solutions may be DEPC<br />

treated and RnaseZap ® or other RNase inhibitors may be used on equipment. Check RNA<br />

quality by agarose gel electrophoresis: 0.5 µg of total RNA should yield two clear bands<br />

of ribosomal RNA (4.5 kb and 1.9 kb).<br />

3. Protocols for smaller amounts of starting material (see Subheadings 2.2., 2.4., 2.6.,<br />

and 3.2.). The <strong>John</strong>s Hopkins protocol as described here (and used personally) requires<br />

a relatively large amount of input material: ideally, at least 2.5 µg of mRNA, broadly<br />

equivalent to 250 µg of total RNA, 250 mg tissue, or 2.5 × 10 7 cultured cells. This<br />

protocol therefore cannot be used to generate expression profiles where RNA is limited,<br />

for example, tissue biopsies. Various technical modifications now enable SAGE to be<br />

applied to smaller quantities of RNA (3): at least 100-fold (and up to 5000-fold) less may<br />

be needed. SADE (a SAGE Adaptation for Downsized Extracts) (6) uses Dynal’s oligo<br />

dT-coated magnetic beads to capture polyA+ mRNA directly from the total RNA or cell<br />

lysate. (This substitutes for mRNA purification then cDNA synthesis with biotinylated<br />

oligo dT followed by capture onto streptavidin-coated Dynabeads.) All steps from<br />

mRNA isolation to tag release are then performed directly on the beads. This procedure<br />

significantly reduces sample loss and has been adopted in Invitrogen’s I-SAGEkit. Oligo<br />

dT is used to similar effect in microSAGE (7) and miniSAGE (8), in the form of a coating<br />

inside microcentrifuge tubes (Roche’s Streptavidin-Coated Tubes). The references contain<br />

the experimental protocols. Further modifications include additional PCR steps. In SADE<br />

and microSAGE, the ditags generated by the first round of large-scale PCR amplification<br />

are then re-amplified using extra PCR cycles (6,7). In contrast, SAGE-Lite (9) and PCR-<br />

SAGE (10) have adapted Clontech’s SMART system to generate PCR-amplified cDNA,<br />

to increase the amount of input material before proceeding to SAGE proper.<br />

4. Enzymatic kinasing of linker oligonucleotides. Dilute linkers to 350 ng/µL. Set up two<br />

tubes, one for linker pair 1 and the other for pair 2. Mix 9 µL of Linker B (either 1B or 2B)<br />

with 8 µL of LoTE, 2 µL of 10× ligase buffer (works with kinase and contains ATP) and<br />

1 µL of T4 polynucleotide kinase. Incubate at 37°C for 30 min then heat inactivate at 65°C<br />

for 10 min. Add 9 µL of Linker 1A to the 20 µL of kinased Linker 1B, and do likewise for<br />

Linkers 2. To anneal linkers, heat to 95°C for 2 min, then place at 65°C for 10 min, 37°C<br />

for 10 min, and room temp for 20 min. Store at –20°C. The final linker concentration is<br />

200 ng/µL. Test the kinase reaction by self-ligating 200 ng of each linker pair. Run on a 12%<br />

polyacrylamide gel. Kinased linkers should result in linker-linker dimers of 80 to 100 bp,<br />

whereas unkinased linkers should not self-ligate. Only linker pairs resulting in over 70%<br />

self-ligation should be used in further steps.<br />

5. Phenolchloroform extraction. If the sample volume is below 200 µL, increase it to 200 µL<br />

with LoTE in a 1.5-mL microcentrifuge tube. Add an equal volume of P/C. Vortex and centrifuge<br />

at full speed for 5 min at room temp. Transfer the aqueous (top) phase to another tube.<br />

6. Ethanol precipitation. For a 200-µL sample volume in a 1.5-mL tube, add 3 µL glycogen<br />

(as a carrier), 100 µL 10 M ammonium acetate, and 700 µL 100% ethanol. (Scale volumes<br />

up or down as required.) Mix well. Precipitate on dry ice (or at –20°C) for at least 15 min.<br />

Centrifuge at full speed for 15 min, preferably at 4°C. Wash pellet with 70% ethanol and<br />

respin. Resuspend in LoTE.<br />

7. QIAquick columns (see Subheading 2.3.). In this protocol, DNA is prepared by P/C extraction<br />

and ethanol precipitation. At some stages, QIAquick ® columns can be substituted,<br />

saving time and possibly providing purer samples for SAGE (12).


282 Oien<br />

8. PCR. With SAGE, achieving the initial 102-bp ditag PCR product takes time. Thereafter,<br />

however, equally strenuous efforts are necessary to avoid PCR cross-contamination, hence<br />

the inclusion of negative controls. The <strong>John</strong>s Hopkins protocol now recommends three<br />

separate PCR areas, for pre-PCR assembly of reagents and negative controls; the addition<br />

of ligated ditag template; and the manipulation of post-PCR products. Ultraviolet PCR<br />

preparation hoods may also be useful.<br />

9. Polyacrylamide gel electrophoresis (PAGE). The SAGE PCR products and ditags are isolated<br />

by 12% PAGE and the concatemers are separated by 8% PAGE. 12% PAGE uses 14 mL<br />

of 40% polyacrylamide (191 acrylamidebis) and 31.3 mL of dH 2 0. 8% PAGE requires<br />

9.3 mL of 40% Polyacrylamide (37.51 acrylamidebis) and 36 mL of dH 2 0. To either,<br />

add 930 µL of 50× Tris acetate buffer, 470 µL of 10% ammonium persulfate, and 30 µL of<br />

TEMED. Mix and pour in a vertical gel apparatus. Allow to polymerize for 30 min. Run<br />

gel as described in the text (the time is approximate). (Note that bromophenol blue is dark<br />

blue and runs ahead of xylene cyanol, which is turquoise.) After electrophoresis, stain gel<br />

with SYBR ® Green I, according to manufacturer’s instructions, or with ethidium bromide.<br />

Visualize bands under ultraviolet light. The <strong>John</strong>s Hopkins and I-SAGE protocols and some<br />

of the methodological papers contain useful photographs of sample gels at each stage.<br />

10. DNA quantitation. The <strong>John</strong>s Hopkins and I-SAGE protocols recommend assessment<br />

of DNA yield by dot quantitation at various stages during PCR purification and ditag<br />

isolation. In general, however, this is not routinely required.<br />

Acknowledgments<br />

Thanks to the Cancer Research Campaign and the University of Glasgow for research<br />

funding and to Dr. Kenneth W. Kinzler for SAGE protocols, software and advice.<br />

References<br />

1. Velculescu, V. E., Zhang, L., Vogelstein, B., and Kinzler, K. W. (1995) Serial analysis of<br />

gene expression. Science 270, 484– 487.<br />

2. Zhang, L., Zhou, W., Velculescu, V., Kern, S., Hruban, R., Hamilton, S., et al. (1997) Gene<br />

expression profiles in normal and cancer cells. Science 276, 1268–1272.<br />

3. Velculescu, V. E., Vogelstein, B., and Kinzler, K. W. (2000) Analysing uncharted transcriptomes<br />

with SAGE. Trends Genet. 16, 423– 425.<br />

4. Madden, S. L., Wang, C. J., and Landes, G. (2000) Serial analysis of gene expression: from<br />

gene discovery to target identification. Drug Discov. Today 5, 415– 425.<br />

5. Polyak, K. and Riggins, G. J. (2001) Gene discovery using the serial analysis of gene<br />

expression technique: implications for cancer research. J. Clin. Oncol. 19, 2948–58.<br />

6. Virlon, B., Cheval, L., Buhler, J. M., Billon, E., Doucet, A., and Elalouf, J. M. (1999) Serial<br />

microanalysis of renal transcriptomes. Proc. Natl. Acad. Sci. USA 96, 15,286–15,291.<br />

7. Datson, N. A., van der Perk-de Jong, J., van den Berg, M. P., de Kloet, E. R., and<br />

Vreugdenhil, E. (1999) MicroSAGE: a modified procedure for serial analysis of gene<br />

expression in limited amounts of tissue. Nucleic Acids Res. 27, 1300–1307.<br />

8. Ye, S. Q., Zhang, L. Q., Zheng, F., Virgil, D., and Kwiterovich, P. O. (2000) MiniSAGE:<br />

Gene expression profiling using serial analysis of gene expression from 1 µg total RNA.<br />

Anal. <strong>Bio</strong>chem. 287, 144–152.<br />

9. Peters, D. G., Kassam, A. B., Yonas, H., O’Hare, E. H., Ferrell, R. E., and Brufsky, A. M.<br />

(1999) Comprehensive transcript analysis in small quantities of mRNA by SAGE-lite.<br />

Nucleic Acids Res. 27, e39.<br />

10. Neilson, L., Andalibi, A., Kang, D., Coutifaris, C., Strauss, J. F., Stanton, J. A. L.,<br />

et al. (2000) Molecular phenotype of the human oocyte by PCR-SAGE. Genomics 63,<br />

13–24.


Serial Analysis of Gene Expression 283<br />

11. Powell, J. (1998) Enhanced concatemer cloning—a modification to the SAGE (Serial<br />

Analysis of Gene Expression) technique. Nucleic Acids Res. 26, 3445–3446.<br />

12. Angelastro, J. M., Klimaschewski, L. P., and Vitolo, O. V. (2000) Improved NlaIII digestion<br />

of PAGE-purified 102 bp ditags by addition of a single purification step in both the SAGE<br />

and microSAGE protocols. Nucleic Acids Res. 28, E62.<br />

13. Margulies, E. H., Kardia, S. L., and Innis, J. W. (2001) Identification and prevention of a<br />

GC content bias in SAGE libraries. Nucleic Acids Res. 29, E60.<br />

14. Kenzelmann, M. and Muhlemann, K. (1999) Substantially enhanced cloning efficiency<br />

of SAGE (Serial Analysis of Gene Expression) by adding a heating step to the original<br />

protocol. Nucleic Acids Res. 27, 917–918.<br />

15. Yamamoto, M., Wakatsuki, T., Hada, A., and Ryo, A. (2001) Use of serial analysis of gene<br />

expression (SAGE) technology. J. Immunol. Methods 250, 45–66.<br />

16. Margulies, E. H. and Innis, J. W. (2000) eSAGE: managing and analysing data generated<br />

with Serial Analysis of Gene Expression (SAGE). <strong>Bio</strong><strong>info</strong>rmatics 16, 650–651.<br />

17. van Kampen, A. H. C., van Schaik, B. D. C., Pauws, E., Michiels, E. M. C., Ruijter, J. M.,<br />

Caron, H. N., et al. (2000) USAGE: a web-based approach towards the analysis of SAGE<br />

data. <strong>Bio</strong><strong>info</strong>rmatics 16, 899–905.<br />

18. Larsson, M., Stahl, S., Uhlen, M., and Wennborg, A. (2000) Expression profile viewer<br />

(ExProView): A software tool for transcriptome analysis. Genomics 63, 341–353.<br />

19. Wheeler, D. L., Church, D. M., Lash, A. E., Leipe, D. D., Madden, T. L., Pontius, J. U.,<br />

et al. (2001) Database resources of the National Center for <strong>Bio</strong>technology Information.<br />

Nucleic Acids Res. 29, 11–16.<br />

20. Lal, A., Lash, A. E., Altschul, S. F., Velculescu, V., Zhang, L., McLendon, R. E., et al. (1999)<br />

A public database for gene expression in human cancers. Cancer Res. 59, 5403–5407.<br />

21. Audic, S. and Claverie, J.-M. (1997) The significance of digital gene expression profiles.<br />

Genome Res. 7, 986–995.<br />

22. Stollberg, J., Urschitz, J., Urban, Z., and Boyd, C. D. (2000) A quantitative evaluation of<br />

SAGE. Genome Res. 10, 1241–1248.<br />

23. Man, M. Z., Wang, X. N., and Wang, Y. X. (2000) POWER_SAGE: Comparing statistical<br />

tests for SAGE experiments. <strong>Bio</strong><strong>info</strong>rmatics 16, 953–959.<br />

24. Kal, A. J., van Zonneveld, A. J., Benes, V., van den Berg, M., Koerkamp, M. G., Albermann,<br />

K., et al. (1999) Dynamics of gene expression revealed by comparison of serial analysis<br />

of gene expression transcript profiles from yeast grown on two different carbon sources.<br />

Mol. <strong>Bio</strong>l. Cell. 10, 1859–1872.<br />

25. Ashburner, M., Ball, C. A., Blake, J. A., Botstein, D., Butler, H., Cherry, J. M., et al.<br />

(2000) Gene ontology: tool for the unification of biology. The Gene Ontology Consortium.<br />

Nat. Genet. 25, 25–29.<br />

26. van den Berg, A., van der Leij, J., and Poppema, S. (1999) Serial analysis of gene expression:<br />

rapid RT-PCR analysis of unknown SAGE tags. Nucleic Acids Res. 27, e17.<br />

27. Chen, J. J., Rowley, J. D., and Wang, S. M. (2000) Generation of longer cDNA fragments<br />

from serial analysis of gene expression tags for gene identification. Proc. Natl. Acad.<br />

Sci. USA 97, 349–353.<br />

28. Matsumura, H., Nirasawa, S., and Terauchi, R. (1999) Transcript profiling in rice (Oryza<br />

sativa L.) seedlings using serial analysis of gene expression (SAGE). Plant J. 20, 719–726.<br />

29. Sambrook, J., Fritsch, E. F., and Maniatis, T. (1987) Molecular Cloning. Cold Spring<br />

Harbor Laboratory Press, Cold Spring Harbor, NY.


284 Oien


Mutation and Polymorphism Detection 287<br />

41<br />

Mutation and Polymorphism Detection<br />

A Technical Overview<br />

Joanne Edwards and <strong>John</strong> M. S. <strong>Bartlett</strong><br />

1. Introduction<br />

Analysis of DNA variation (polymorphism and mutations) is one of the most<br />

common challenges faced by molecular biologists. Studies of polymorphisms and<br />

mutations as molecular markers of or underlying causes of disease have confirmed<br />

the importance of mutation and polymorphism detection. With mutation detection<br />

currently being so important for the study of genetic diseases, gene discovery, and<br />

solving problems of basic biology, there is a large demand for quick and relatively<br />

cheap methods for mutation detection. Therefore, many different methods have been<br />

developed for detecting new mutations and screening populations for known mutations<br />

or polymorphisms. Traditional mutation detection systems, such as restriction fragment<br />

length polymorphism and denaturing gradient gel electrophoresis, have their limitations.<br />

With restriction fragment length polymorphism, the mutational event needs to either<br />

create or destroy a restriction site (1). With denaturing gradient gel electrophoresis,<br />

although a change at a restriction digest site is not required, this method may only<br />

detect about 50% of possible mutations and polymorphisms (1). The use of polymerase<br />

chain reaction (PCR)-based mutation and polymorphism detection systems<br />

increase the sensitivity and accuracy of the screening methods. This chapter will introduce<br />

some of the available detection methods and discuss the problems associated<br />

with these techniques.<br />

2. Detection Methods<br />

2.1. Detection of Single Nucleotide Polymorphisms (SNPs)<br />

2.1.1. Single-Strand Conformation Polymorphism (SSCP)<br />

SSCP can detect up to 90% of single base changes and relies on the different<br />

mobilities of DNA strands containing single bp differences when run on a denaturing<br />

polyacrylamide gel (2). SSCP analysis is widely used to screen large numbers of samples<br />

for mutations. If a mutation is detected, direct sequence analysis is often then used to<br />

determine the exact location and base change of the mutation. SSCP can be performed<br />

using either fluorescent or radioactive technology (2,3). The advantages of fluorescent<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

287


288 Edwards and <strong>Bartlett</strong><br />

technology over radioactive are (1) an internal molecular weight standard is used in each<br />

lane to align data, thereby eliminating lane-to-lane variability; (2) strands of DNA in the<br />

PCR amplicon can be labeled with different-colored labels; And (3) there is the ability<br />

to multiplex products from different PCRs in one lane to increase sample throughput.<br />

SSCP is further discussed in Chapter 48. However, although SSCP is a useful screening<br />

technique, it yields no <strong>info</strong>rmation on the location of the mutation. A band shift will<br />

identify the presence of the mutation within the PCR product, but no sequence <strong>info</strong>rmation<br />

on the exact base that has been mutated can be determined without sequencing.<br />

2.1.2. Mutational Detection Using Cleavage Systems<br />

2.1.2.1. GEL-BASED METHOD<br />

An alternative screening method for mutations or polymorphisms that does provide<br />

some <strong>info</strong>rmation on the location of the mutation is enzymatic cleavage of DNA.<br />

Enzymatic cleavage systems rely on enzymes such as resolvases (e.g., the enzyme T4<br />

endonuclease VII; ref. 4). These enzymes cleave double-stranded DNA at sites where<br />

a “bubble” is formed because of miss-pairing of bases. Mutation analysis is performed<br />

by mixing PCR products from the test samples and samples with known sequences<br />

(either normal alleles or wild type DNA). The PCR products are melted and the DNA<br />

strands allowed to re-anneal. In the presence of either a mutation or a different allele<br />

from the wild type, a miss-match in sequence provides a target for the resolvase<br />

enzyme. The enzyme scans along the double-stranded DNA and binds to the singlestranded<br />

“bubble” at the site of the miss-match, and the DNA is then cleaved in<br />

this region. The resultant fragments are then separated by electrophoresis, and the<br />

presence of cleavage products indicates the presence of a mutation (4,5). The size of<br />

the cleaved fragments indicates the approximate location of the mutation, which is<br />

an advantage over SSCP.<br />

2.1.2.2. NONGEL-BASED METHOD<br />

The method just discussed uses a gel-based detection system. An example of a<br />

cleavage method that is not gel based is the use of a structure-specific 5′ nuclease to<br />

cleave sequence-specific structures in each of two cascading reactions (6). The cleavage<br />

structure forms when two synthetic oligonucleotides hybridize in tandem to a target.<br />

One of the oligonucleotides cycles on and off the target and is cut by the nuclease<br />

only when the appropriate structure forms. The cleaved probes then participate in a<br />

second reaction involving a dye-labeled fluorescence resonance energy-transfer probe.<br />

Cleavage of this probe generates a signal, which can be analyzed by fluorescence<br />

microtiter plate readers (6).<br />

2.1.3. Oligonucleotide Microarrays<br />

Current screening methods are rapid and when combined with gel-based fluorescent<br />

DNA sequencing technology can accurately locate and identify mutations and polymorphisms.<br />

However, SNPs are now being uncovered and assembled into large SNP<br />

databases. Although this will be invaluable for linking SNPs with disease, it will require<br />

analysis of thousands of polymorphisms (7). Currently, there is no technique available<br />

that is fast, specific, sensitive, reliable, and cost-effective enough for genome-wide<br />

polymorphism analysis. Oligonucleotide microarrays are currently under intensive


Mutation and Polymorphism Detection 289<br />

development and if this technique is successful should enable the investigation of<br />

thousands of SNPs in parallel (8). However, currently there are still limitations<br />

associated with it. A recent mapping study analyzed 500 SNPs using array technology<br />

and only 70% of sites analyzed were genotyped correctly (9). This limitation has been<br />

broached by using enzymatic reactions with the oligonucleotide arrays to enhance<br />

specificity of hybridization, for example, array primer extensions are being developed<br />

that use the principle of allele specific single-base primer extension. It has been<br />

demonstrated that this method is more specific at identifying unknown mutations<br />

in a known sequence (deletions, transversions, and up to two-base insertions) that<br />

can be readily found, using arrayed primer extension reactions (10). However,<br />

promising mutational analysis using oligonucleotide microarrays still remains in the<br />

developmental stage.<br />

2.2. Detection of Large Inserts or Deletions<br />

Large inserts or deletions may be detected by the methods discussed above, such as<br />

SSCP, enzyme cleavage systems, or amplification refractory mutation system discussed<br />

in detail in Chapter 47. However, currently the most commonly used detection system<br />

is either radioactive or fluorescent-based automated sequencing. Although many of<br />

the above mentioned available screening methods detect in the region of 90 to 98%<br />

of mutations (whether SNPs, deletions, or insertions) there are still mutations that are<br />

missed. Therefore, currently the only way to assure that every mutation is found is to<br />

sequence the region of interest, although this method is both costly and time consuming<br />

and requires high-quality pure DNA.<br />

2.3. Detection of Loss of Heterozygosity (LOH)<br />

and Replication Error Phenotype<br />

Microsatellite loci have a high degree of polymorphism that is caused by problems<br />

associated with copying repetitive sequences of DNA. Microsatellites are popular<br />

genetic markers because of their abundance and high level of allelic variation, and<br />

expansion of microsatellite trinucleotide repeats have been demonstrated to cause<br />

several human genetic disease (11). Microsatellite analysis is also useful for detection<br />

of LOH and replication errors in tumors (12,13). LOH is detected as a reduced intensity<br />

or total loss of one or more bands in the tumor DNA compared with normal DNA<br />

form the same individual (12). Replication errors are detected as changes in the length<br />

of the microsatellite sequences in the tumor DNA compared with normal DNA (13).<br />

Mutation polymorphism detection of microsatellites basically involves amplifying<br />

the microsatellite loci and estimating the size of the PCR product. The most common<br />

techniques used to estimate the size of the microsatellite product involves separation<br />

of products by gel electrophoresis. The most common method used to visualize PCR<br />

products on an agarose gel is ethidium bromide staining. Staining with ethidium<br />

bromide is not a routine method used for microsatellite analysis because it has low<br />

sensitivity and does not provide a permanent record. Microsatellite analysis is more<br />

commonly performed using polyacrylamide gels because they give better separation<br />

than agarose gels. Visualization of DNA on a polyacrylamide gel may be achieved<br />

by silver staining, radioactive-labeled DNA visualized by autoradiography, or laseractivated<br />

fluorescent-labeled DNA detected by an automated sequencer. Silver staining<br />

can be difficult to control, and use of radiation within the laboratory setting is becom-


290 Edwards and <strong>Bartlett</strong><br />

ing less popular in current years because of the associated health risks. Therefore,<br />

fluorescent-based methods are currently the method of choice, that is, labeling of DNA<br />

using fluorescent-labeled primers or nucleotides. Fluorescent technology eliminates the<br />

handling of radioactivity, makes it easier to interpret stutter bands and laser-activated<br />

detection of fluorescent products during electrophoresis, allows immediate detection of<br />

signal, permits rapid data analysis, and provides permanent records of results. The use<br />

of fluorescent technology has allowed development of techniques that greatly increase<br />

the speed by which microsatellite analysis can be performed by increasing throughput.<br />

For example, a rapid way to screen for microsatellites is described in Chapter 42. This<br />

chapter discusses multiplex touchdown PCR, which allows amplification of multiple<br />

microsatellite loci simultaneously. Use of different fluorescent-labeled primers then<br />

allows identification of each loci. Similarly, Chapter 43 describes the use of fluorescent<br />

technology to increase throughput of LOH analysis by labeling PCR products with<br />

differently labeled fluorescent primers.<br />

3. Technical Constraints<br />

All of the above methods require good-quality DNA; therefore, in most cases before<br />

mutational analysis can begin RNA or DNA must be extracted from the sample and a PCR<br />

performed to provide DNA of sufficient purity and quantity for mutational analysis.<br />

3.1. Problems Associated with Quality and Quantity<br />

of Extracted DNA or RNA<br />

If the amount of available DNA or RNA is limited and therefore insufficient for<br />

use in mutational analysis, PCR or RT-PCR may be used to increase the quantity of<br />

DNA. If multiple regions of interest are to be investigated, then it may be necessary to<br />

globally amplify DNA by degenerate oligonucleotide-primed PCR (DOP-PCR) (14).<br />

The use of this technique in mutational analysis is described in Chapter 45.<br />

The purity of extracted DNA or RNA for subsequent mutational analysis is extremely<br />

important. Contamination with other cellular components, such as protein, or chemicals,<br />

such as ethanol, can easily be removed after the extraction process using various<br />

cleanup protocols. However, contamination with foreign DNA is not dealt with as<br />

easily. Contamination at this stage of the process with either foreign DNA from external<br />

sources or from within the sample will be amplified as the process proceeds and<br />

can result in mutational analysis being conducted on the wrong DNA. Precautions<br />

should be taken to limit cross-contamination and external contamination as with any<br />

PCR. Methods for increasing the purity of RNA and decreasing the introduction of<br />

contamination during RT-PCR are discussed in Chapter 46. Contamination may also be<br />

a problem from within the tissue, for example, if studying cancer genetics, normal cells<br />

may contaminate a tumor cell population. Microdissection is a technique that allows<br />

separation of normal and tumor cells, and Chapter 43 discusses use of this method<br />

in microsatellite analysis.<br />

For mutational analysis, DNA is required to be of high quality because poor-quality<br />

DNA, for example, degraded or nicked, could result in false results. If performing<br />

RT-PCR from RNA, RNA should be extracted from fresh tissue or from tissue snap<br />

frozen on removal to decrease RNA degradation. Although kits now are available for<br />

extraction of RNA from archival material, the quality of RNA retrieved is often poor.


Mutation and Polymorphism Detection 291<br />

DNA may be extracted from fresh, frozen, or archival material. High-quality DNA<br />

should be easily retrieved from fresh or frozen material, for example, tissue or blood.<br />

Extraction of good-quality DNA from archival material is more difficult; however, the<br />

most abundant source of clinically available tissue is archival formalin-fixed paraffinembedded<br />

tissue. This process not only drastically decreases the yield of DNA in<br />

comparison with fresh or frozen material but also damages the DNA, often making it<br />

unsuitable as a DNA template. Chapter 43 discusses how archival DNA can be used for<br />

microsatellite analysis by keeping the PCR product size small.<br />

In summary, for successful mutational analysis a significant quantity of high-quality,<br />

pure DNA is required. Low quantities of DNA make analysis difficult, and poor-quality<br />

or contaminated DNA may cause false results.<br />

3.2. PCR Artifacts<br />

Although PCR is the recognized method for increasing the quantity of DNA required<br />

for polymorphism or mutational analysis, PCR can in itself introduce problems.<br />

Contamination by foreign DNA and cross-contamination is often introduced at this<br />

stage of the process. With appropriate care, that is, gloves, aseptic technique, and<br />

clean pipette tips between every sample, this should be limited. However, this is<br />

a particular problem with DOP-PCR. DOP-PCR is very sensitive to contamination<br />

because degenerate primers in the reaction will amplify DNA from any source present<br />

in the tube. However, the presence of DNA in the blank control after DOP-PCR does<br />

not always indicate contamination, and the low temperature in the DOP-PCR can result<br />

in DOP primers attaching to each other and replicating (15). If this occurs, then a<br />

second PCR amplifying only the region of interest should not be possible. It is therefore<br />

recommended that blanks should be tested by sequential PCR after DOP-PCR before<br />

samples are discarded.<br />

Other problems created by PCR are the addition of an extra base (normally adenine) at<br />

the end of the DNA sequence and the creation of shadow bands (discussed in Chapter 44).<br />

In summary, although PCR solves the problem of insufficient DNA for analysis, with<br />

its use comes an assortment of additional problems.<br />

4.1. Applications<br />

Mutational analysis has many applications, for example, markers of disease, cancer<br />

research, and genotyping. Mutations are responsible for diseases, such as sickle cell<br />

anemia (1), cystic fibrosis (1), and adrenal leukodystrophy (1). An expansion of the<br />

CAG repeat in exon 1 of the androgen receptor alone can result in Kennedy’s disease,<br />

spinocerebellar ataxia, or fragile X syndrome (15). Therefore, polymorphism and<br />

mutational analysis are of use in diagnosing these diseases.<br />

Detection of polymorphism and mutational analysis is also widely used in the field<br />

of cancer research. Mutation detection has demonstrated that p53 function is lost in<br />

approximately 50% of all cancers and that the loss of function is caused by point<br />

mutations (16). It also demonstrates that a mutation in BRCA 1 or BRCA 2 increases<br />

susceptibility to breast and ovarian cancer (17). Microsatellite analysis is used in cancer<br />

research to study LOH and microsatellite instability. LOH studies have demonstrated<br />

most genetic alterations that occur in bladder cancer are on chromosome 9 (18) and that<br />

microsatellite instability is commonly found in colorectal cancer (13).


292 Edwards and <strong>Bartlett</strong><br />

Stepwise mutation models and microsatellite polymorphisms have been used to<br />

determine relationships among primates (19). Such studies have investigated the<br />

relationship between humans, chimpanzees, and gorillas (19). There is also an ongoing<br />

debate about proper interpretation of DNA sequence polymorphisms and their ability<br />

to reconstruct human population history (20).<br />

Mutational analysis is also used in genotyping. It was use of mutational analysis<br />

that demonstrated that birds are not monogamous because genotyping performed on<br />

eggs in the same nest demonstrated different parentage (21). Genotyping is also used<br />

when breeding animals in captivity to determine their most appropriate mate (21).<br />

In summary, polymorphism and mutational analysis is a major tool in science today,<br />

without which much of our current understanding of genetics would not have been<br />

possible.<br />

References<br />

1. Alberts, B., Bray, D., Lewis, J., Raff, M., Roberts, K., and Watson, J. D. (1989) Recombinant<br />

DNA Technology, in Molecular <strong>Bio</strong>logy of The Cell, 2nd ed., chapter 4, (Robertson, M., ed.),<br />

Garland Publications, New York, pp. 180–196.<br />

2. Iwahana, H., Yoshimoto, K., and Itakura, J. (1994) Multiple fluorescence-based PCR-SSCP<br />

analysis. <strong>Bio</strong>techniques 16, 296–304.<br />

3. Nataraj, A. J., Olivos-Glander, I., Kusukawa, N., and Highsmith, W. E. (1999) Single-strand<br />

conformation polymorphism and heteroduplex analysis for gel-based mutation detection.<br />

Electrophoresis 20, 1177–1185.<br />

4. Youil, R., Kemper, B. W., and Cotton, R. G. (1995) Screening for mutations by enzyme<br />

mismatch cleavage with T4 endonuclease VII. Proc. Natl. Acad. Sci. USA 92, 87–92.<br />

5. Del Tito, B. J., Poff, H. E., Noviotny, M. A., Cartlidge, D. M., Walker, R. I., Earl, C. D.,<br />

et al. (1998) Automated fluorescent analysis procedure for enzymatic mutation detection.<br />

Clin. Chem. 44, 731–739.<br />

6. Cotton, R. G. (1999) Mutation detection by chemical cleavage. Genet. Anal. 14, 165–168.<br />

7. Gray, I. C., Campbell, D. A., and Spurr, N. K. (2000) Single nucleotide polymorphisms as<br />

tools in human genetics. Hum. Mol. Genet. 9, 2403–2408.<br />

8. Kikoris, M., Dix, K., Moynihan, K., Mathis, J., Erwin, B., Grass, P., et al. (2000) High<br />

throughput SNP genotyping with the mass code system. Mol. Diagn. 5, 329–340.<br />

9. Faham, M., Baharloo, S., Tomitaka, S., DeYoung, J., and Freimer, N. B. (2001) Mismatch<br />

repair detection: high-throughput scanning for DNA variations. Hum. Mol. Genet. 10,<br />

1657–1664.<br />

10. Fan, J. B., Chen, X., Halushka, M. K. Berno, A., Huang, X., Ryder, T., et al. (2000) Parallel<br />

genotyping of human SNPs using generic high-density oligonucleotide tag arrays. Genome<br />

Res. 10, 853–860.<br />

11. Rohrbach, H., Hass, C .J., Baretton, G. B., Hirschmann, A., Diebold, J., Behrendt,<br />

R. P., et al. (1999) Microsatellite instability and loss of heterozygosity in prostatic<br />

carcinomas: comparison of primary tumours, and of corresponding recurrences after<br />

androgen-deprivation therapy and lymph node metastases. Prostate 40, 20–27.<br />

12. Niederacher, D., Picard, F., van Roeyen, C., An, A. X., Bender, H. G., and Beckmann,<br />

M. W. (1997) Patterns of Allelic loss on chromosome 17 in sporadic breast carcinomas<br />

detected by fluorescent labelled microsatellite analysis. Genes Chromosomes Cancer 18,<br />

181–192.<br />

13. Dietmaier, W., Riedlinger, W., Kohler, A., Wegele, P., Beyser K., Sagner, G., et al. (1999)<br />

Detection of microsatellite instability and loss of heterozygosity in colorectal tumours by<br />

fluorescence based multiplex microsatellite PCR. <strong>Bio</strong>chemica 2, 43– 46.


Mutation and Polymorphism Detection 293<br />

14. Kuukasjarvi, T., Tanner, M., Pennanen, S., Karhu, R., Visakorpi, T., and Isola, J. (1997)<br />

Optimising DOP-PCR for universal amplification of small DNA samples in comparative<br />

genomic hybridisation. Genes, Chromosomes Cancer 18, 94–101.<br />

15. Chamberlain, N. L., Driver, E. D., and Mainsfeld, R. L. (1994) The length and location<br />

of CAG trinucleotide repeats in the androgen receptor N-terminal domain affect trans<br />

activation function. Nucleic Acids Res. 22, 3181–31886.<br />

16. Lane, D. P. (1992) TP53, the guardian of the genome. Nature 358, 15–16.<br />

17. Rohlfs, E. M., Learning, W. G. Friedman, K. J., Cough, F. J., Weber, B..L., and Silverman,<br />

L. M. (1997) Direct detection of mutations in the breast and ovarian cancer susceptibility<br />

gene BRCA1 by PCR mediated site directed mutagenesis. Clin. Chem. 43, 24–29.<br />

18. Knowles, M. A. (1998) Molecular genetics of bladder cancer: Pathways of development<br />

and progression. Cancer Surveys 31, 49–53.<br />

19. Goldstein, D. B., Ruis Linares, A., Cavalli-Sforsa, L. L., and Feldman, M. W. (1995)<br />

Genetic absolute dating based on microsatellite and the origin of modern humans. Proc.<br />

Natl. Acad. Sci. USA 92, 6723–6727.<br />

20. Cann, R. L. (2001) Genetic clues to dispersal in human populations: retracing the past from<br />

the present. Science, 291, 1742–1748.<br />

21. Alberts, B., Bray, D., Lewis, J., Raff, M., Roberts, K., and Watson, J. D. (1989) The<br />

organisation and evolution of the nuclear genome, in Molecular <strong>Bio</strong>logy of The Cell,<br />

2nd ed, chapter 10 (Robertson, M., ed.), Garland Publications, New York, pp. 599–609.


294 Edwards and <strong>Bartlett</strong>


MT-PCR for Microsatellite Analysis 295<br />

42<br />

Combining Multiplex and Touchdown PCR<br />

for Microsatellite Analysis<br />

Kanokporn Rithidech and <strong>John</strong> J. Dunn<br />

1. Introduction<br />

An improved nonradioactive polymerase chain reaction (PCR)-based method for<br />

simultaneous amplification of multiple loci of microsatellites has been developed as a<br />

rapid way to screen for microsatellites (1). The approach, termed multiplex-touchdown<br />

PCR (MT-PCR), is performed in a single PCR tube by combining touchdown (2–5) and<br />

multiplex (6) PCR protocols. The touchdown format is used to improve the specificity<br />

and the quality of amplification, that is, only DNA bands of an expected size are<br />

present as the major PCR bands observed on the nondenaturing polyacrylamide gels,<br />

thereby overcoming the presence of differently sized background bands (a ladder-like<br />

problem). The multiplex strategy is used so that simultaneous amplification of multiple<br />

microsatellite loci is achieved. In this chapter, we describe the MT-PCR strategy<br />

that has been successfully used for simultaneous amplification of up to three mouse<br />

microsatellites by choosing primer pairs with the corresponding touchdown-PCR<br />

parameters. The MT-PCR is very useful for genotyping hybrid mice, provided the<br />

allelic size difference between two parental genotypes is amenable to separation by<br />

gel electrophoresis. In principle, this MT-PCR should be applicable to similar studies<br />

in other species, including humans.<br />

2. Materials<br />

1. Sterile deionized water, for example, MilliQ water.<br />

2. 10× PCR buffer (10 mM Tris-HCL, pH 8.3, 50 mM KCL).<br />

3. 10 mM of each of dNTP (A,T,C, and G).<br />

4. Microsatellite primers.<br />

5. Taq polymerase.<br />

6. MgCl 2 .<br />

7. Template DNA (5 µg/mL, in sterile water).<br />

8. Polyacrylamide (premixed 300.8 acrylamide:bisacrylamide, Owl Separation System<br />

Portsmouth, NH).<br />

9. Slab gel, for example, Mini-protean cell from <strong>Bio</strong>-Rad, large format single sided vertical<br />

system (20 cm × 20 cm) from Owl separation System.<br />

10. 1× TBE (Tris borate EDTA) gel running buffer.<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

295


296 Rithidech and Dunn<br />

11. Ethidium bromide (0.5 µg/mL).<br />

12. DNA size standards.<br />

3. Methods<br />

Basically, two basic principles are involved in this protocol, that is, 1) the repeat<br />

sequences are amplified by PCR by using primers specific for the flanking genes and<br />

2) the amplicons are then sized on nondenaturing polyacrylamide gels. The protocol<br />

outlined below has been used successfully for genotyping 50 mouse microsatellite<br />

markers, included in our study on leukemogenesis and hepatoma, of BALB/cJ X<br />

CBA/CaJ hybrid mice. The sizes of these markers are 84 to 270 bp with allelic size<br />

differences of 8 to 30 bp. The following steps are required.<br />

3.1. Step 1<br />

Set up PCR Master Mix. An example given in below is for a single PCR amplification<br />

of three microsatellite markers. To make a total volume of 15 µL Master Mix, combine<br />

the appropriate set of components in order listed in below in a 0.5-mL thin-walled<br />

PCR tube.<br />

Volume (µL)<br />

Final Concentration<br />

Sterile Milli Q Water 1.19<br />

10× PCR buffer 1.5 1X<br />

dATP at 10 mM 0.3 200 µM<br />

dCTP at 10 mM 0.3 200 µM<br />

dGTP at 10 mM 0.3 200 µM<br />

dTTP at 10 mM 0.3 200 µM<br />

Microsatellite marker 1:<br />

Forward Primer (6.6 µM) 1.14 0.5 µM<br />

Reverse Primer (6.6 µM) 1.14 0.5 UM<br />

Microsatellite marker 2:<br />

Forward Primer (6.6 µM) 1.14 0.5 µM<br />

Reverse Primer (6.6 µM) 1.14 0.5 µM<br />

Microsatellite marker 3:<br />

Forward Primer (6.6 µM) 1.14 0.5 µM<br />

Reverse Primer (6.6 µM) 1.14 0.5 µM<br />

MgCl 2 (25 mM) 1.2 2.0 mM<br />

template DNA (5 µg/mL) 3.0 1.0 µg/mL<br />

Taq polymerase (5Units/µL) 0.072 2.5 Units/100 µL<br />

NOTES:<br />

1. All chemicals, except primers, are purchased from Perkin-Elmer, Norwalk, CT.<br />

2. All primers are purchased from Research Genetics, Inc. Huntsville, AL.<br />

Add template DNA after a drop of mineral oil (Sigma, St. Louis, MO) has been placed<br />

into the tube.<br />

Add after “Hot Start” in Step 2.<br />

3.2. Step 2<br />

Hot-start at 96°C for 3 min. Then, add 0.072 µL of Taq polymerase (5 U/µL).


MT-PCR for Microsatellite Analysis 297<br />

3.3. Step 3: Touchdown<br />

A typical PCR protocol for the initial Touchdown cycle is denaturation at 96°C for<br />

30 s, annealing at 65°C for 30 s, and extension at 72°C for 30 s.<br />

3.4. Step 4: Standard PCR Cycles<br />

After 5 to 10 touchdown cycles (depending on the targeted annealing temperature),<br />

20 standard PCR cycles are performed under the following conditions: denaturation at<br />

96°C for 30 s, annealing at 55°C (or 60°C) for 30 s, extension at 72°C for 30 s, and<br />

final extension at 72°C for an additional 5 min (see Note 3).<br />

3.5. Step 5: Gel Electrophoresis<br />

Typically, 6 and 10% nondenaturing polyacrylamide gels are used for separating<br />

microsatellites with allelic size differences of >12 bp and 20 cycles) results in deterioration of<br />

the PCR specificity leading to an increase in the background of spurious bands. Under<br />

the proper conditions, several amplicons from one round of MT-PCR can be analyzed<br />

simultaneously and each amplicon can be identified unambiguously on the same lane of<br />

nondenaturing PAGE. By choosing primer pairs with the corresponding T-PCR parameters,<br />

the MT-PCR is highly efficient for concurrent identification of three microsatellite loci<br />

(Fig. 1, lane 1).<br />

3. A soak file at 4°C can be set after the final extension step.<br />

4. A Mini-protean cell from <strong>Bio</strong>-Rad is used for the 6% nondenaturing polyacrylamide gel.<br />

A large format single sided vertical system (20 cm × 20 cm) from Owl separation System<br />

is used for the 10% nondenaturing polyacrylamide gel.<br />

5. Beside speed and ease, an additional advantage of MT-PCR is that it is notably useful<br />

when only small quantities of DNA template are available. Amounts of DNA template<br />

as low as 10 ng can be used in one simultaneous reaction with three different pairs of<br />

primers. DNA samples from different tissues of the mouse, that is, bone marrow cells,<br />

spleen, liver, and tail, have been used successfully as templates in MT-PCR. Normally,<br />

the simple salting out procedure for isolating genomic DNA described by Miller et al. (8)<br />

routinely gives, in our laboratory, high-quality DNA templates from these tissues. The<br />

only drawback to the use of MT-PCR for the genotyping of microsatellites is that each<br />

protocol depends largely on primer compatibility, that is, having similar reaction kinetics<br />

(amenable to the same T-PCR protocol).


298 Rithidech and Dunn<br />

Fig. 1. PCR amplification of the mouse microsatellites: D2Mit43 (242-210 bp), D2Mit107<br />

(134-112 bp, and D2Mit126 (160–190 bp) using genomic DNA isolated from the bone marrow<br />

cells of C3HeB/HeJ X C57BL/6J F1 Hybrid. Products were separated in a 6% nondenaturing<br />

PAGE (300.8 acrylamidebisacrylamide) run at 150 V (constant) for 3 h. The gel was stained<br />

with ethidium bromide (0.5 mg/mL) and photographed under ultraviolet light. The arrows<br />

indicate the position of the expected amplicons. The molecular markers (HaeIII digested<br />

pBR332 DNA) are shown in lanes M. The PCR products were obtained using five Touchdown<br />

cycles, and 20 cycles of constant annealing temperature (60°C). Lane 2 represents a control<br />

PCR without added mouse DNA. The sequences of primers are as follows:<br />

Acknowledgments<br />

This work was support by the U.S.DOE Contract ER62487/0001684.<br />

References<br />

1. Rithidech, K., Dunn, J. D., and Gordon, C. R. (1997) Combining multiples and touchdown<br />

PCR to screen murine microsatellite polymorphisms. <strong>Bio</strong>techniques 23, 36– 44.<br />

2. Chou, Q., Russell, M., Birch, D. E., Raymond, J., and Bloch, W. (1992) Prevention of<br />

pre-PCR mis-priming and primer dimerization improves low-copy-number amplifications.<br />

Nucleic Acids Res. 20, 1717–1723.<br />

3. Don, R. H., Cox, P. T., Wainright, B. J., Baker, K., and Mattick, J. S. (1991) Touchdown<br />

PCR to circumvent spurious priming during gene amplification. Nucleic Acids Res. 19,<br />

4008.<br />

4. Scrimshaw, B. J. (1992) A simple nonradioactive procedure for visualization of (dC-dA)n<br />

dinucleotide repeat length polymorphisms. <strong>Bio</strong>Techniques 13, 189.


MT-PCR for Microsatellite Analysis 299<br />

5. Mellersh, C. and Sampson, J. (1993) Simplifying detection of microsatellite length<br />

polymorphisms. <strong>Bio</strong>Techniques 15(4), 582–583.<br />

6. Chamberlain, J. S., Gibbs, R. A., Ranier, J. E., Nguyen, P. N., and Caskey, C. T. (1988)<br />

Detection screening of the Duchenne muscular dystrophy via multiplex DNA amplification.<br />

Nucleic Acids Res. 16, 11,141–11,156.<br />

7. Thein, S. L. and Wallace, R. B. (1986) The use of synthetic oligonucleotides as specific<br />

hybridization probes in the diagnosis of genetic disorders, in Human genetic diseases: a<br />

practical approach (Davis, K. E., ed.), IRL Press, Herndon, Virginia, pp. 33–50.<br />

8. Miller, S. A., Dykes, D. D., and Polesky, H. F. (1989) A simple salting out procedure for<br />

extracting DNA from human nucleated cell. Nucleic Acids Res. 16, 1215.


300 Rithidech and Dunn


Detection of MIS 301<br />

43<br />

Detection of Microsatellite Instability and Loss<br />

of Heterozygosity Using DNA Extracted from<br />

Formalin-Fixed Paraffin-Embedded Tumor Material<br />

by Fluorescence-Based Multiplex Microsatellite PCR<br />

Joanne Edwards and <strong>John</strong> M. S. <strong>Bartlett</strong><br />

1. Introduction<br />

Microsatellites are widely distributed highly repetitive DNA sequences composed of<br />

di-, tri-, or tetranucleotide repeats (1). They are spread over the whole human genome<br />

and are located between and within genes. Physiologically, they exhibit high levels of<br />

polymorphism, relative to different chromosomal loci, and within different individuals,<br />

different microsatellite lengths can even be noted between two alleles of the same<br />

gene (1). Although the role of microsatellites in the genome and their evolutionary<br />

mechanisms are still incompletely understood, they are widely used tools in genetic<br />

mapping studies and as markers for prediction of disease (2,3).<br />

The length of microsatellite repeats may be used to predict the presence of disease,<br />

for example, the length of the CAG repeat in the AR receptor gene (40–65 repeats)<br />

is a marker of Kennedy’s disease (3). Microsatellites also exhibit a form of genetic<br />

instability characterized by the gain or loss of repeat units at multiple independent loci.<br />

Such alterations have been observed to accumulate in cells with defective DNA repair<br />

mechanisms and are commonly known as microsatellite instability (MIS) (1,2). The<br />

DNA mismatch repair system plays an important role in controlling the accumulation of<br />

somatic mutations and therefore there is an association between DNA repair defects and<br />

carcinogenesis, and the presence of MIS may be used as a marker of certain cancers (4).<br />

MIS has been associated with familial cancer syndrome hereditary nonpolyposis colorectal<br />

cancer, sporadic ovarian, prostate. and pancreatic cancer (1,4,5). Cancer research<br />

also uses microsatellites when investigating inactivation of tumor suppressor genes.<br />

Loss of heterozygosity (LOH) of microsatellites located in or close to a tumor suppressor<br />

gene is an indirect way of testing for inactivation of tumor suppressor genes in<br />

accordance to the classical two-hit model, that is, a recessive mutation is uncovered by<br />

loss of the second copy of the gene, resulting in inactivation of the tumor suppressor<br />

gene (6).<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

301


302 Edwards and <strong>Bartlett</strong><br />

Studies that involve microsatellites usually involve amplification of DNA using<br />

polymerase chain reaction (PCR). The three methods for detection of microsatellites<br />

currently in use are: radioactive PCR visualized by autoradiography; PCR visualized<br />

by silver-stained gels; and fluorescent PCR visualized on an automated DNA fragment<br />

analyzer (7). The current method of choice is fluorescent PCR coupled with analysis<br />

of fluorescent-labeled DNA fragments on polyacrylamide gel electrophoresis in an<br />

automated DNA fragment analyzer. This method considerably increases the throughput<br />

of samples because it uses detection of laser-activated fluorescent products during<br />

electrophoresis and data are then immediately available for analysis. A large number<br />

of microsatellites may be studied at once because as many as 20 PCR products may<br />

be run down one lane of the gel at any one time. This method is also of use if there is a<br />

limited amount of DNA template, for example, small biopsy samples, because it enables<br />

the detection of very low concentrations of PCR products. The fluorescence detection<br />

system has been proven to be at least 10 times more sensitive than autoradiography<br />

and silver staining because of the use of an internal standard the fluorescent system<br />

can more accurately size PCR products and it minimizes stutter bands. This method<br />

also enables detection of LOH even if complete allele loss is not apparent because of<br />

contamination from other cell types. There are also less associated hazards when using<br />

this technique because no handling of radioactivity is required.<br />

LOH and MIS studies commonly are used in cancer research; therefore, tumor and<br />

normal tissue must be separated before DNA can be extracted. Microdissection of<br />

histologically characterized cells from fresh-frozen or paraffin-embedded tissue sections<br />

has become an important technique for the analysis of genetic alterations occurring<br />

in tumors, allowing the separation of normal and tumor cells. Microdissection is a<br />

technique where by tissue sections approx 5-µm thick are stained in toluidine blue and<br />

viewed using a light microscope, and areas of normal/tumor cells are dissected using a<br />

micromanipulator. DNA from the different cells is extracted from the dissected tissue.<br />

The most abundant source of tumor tissue with clinical <strong>info</strong>rmation is formalin-fixed<br />

paraffin-embedded tissue. Hospitals routinely archive tumors in this fashion, and<br />

although this is acceptable for most pathological processes, it does cause nicks to<br />

form in DNA, this damage to the DNA affects the ability to use it as a PCR template.<br />

The yield of DNA extracted from formalin fixed paraffin embedded tissue is also<br />

significantly diminished, (by about 70% compared to the yield from frozen tissue) (8).<br />

To overcome the problems associated with using archival DNA as a PCR template, the<br />

PCR product should be kept to a minimum size when designing primers, preferably<br />

under 300 bp. To overcome the concentration problem, fluorescent PCR with 40 or<br />

more amplification cycles should be used. Hot start Taq polymerase should also be used<br />

to decrease nonspecific binding of primers to the DNA template and the formation<br />

of primer dimers. With hot start Taq, the amplification reaction only starts after the<br />

initial denaturation step. Therefore, nonspecific PCR products, primer dimer formation,<br />

and background are minimized, maximizing the yield of specific PCR product from<br />

limited template.<br />

The following method is an example of how MIS and LOH can be determined for a<br />

panel of nine microsatellites spanning chromosome 9 using only 9 µL of DNA (approx<br />

15 µg/µL) microdissected from formalin-fixed paraffin-embedded archival transitional<br />

cell carcinomas of the urinary bladder tissue.


Detection of MIS 303<br />

Table 1<br />

Sequence of the Forward and Reverse Primers for Each Microsatellite<br />

and the Annealing Temperature Used in the PCR<br />

Primer name Primer sequence Annealing temp<br />

D9S126<br />

Forward ATT GAA ACT CTG CTG AAT TTT CTG 50°C<br />

Reverse<br />

CAA CTC CTC TTG GGA ACT GC<br />

D9S259<br />

Forward GGC ATC ATT GCA CCA T 50°C<br />

Reverse<br />

GGA TGG ATC TTA TGG GTG GAA<br />

D9S275<br />

Forward CAG GAA CTT GTC CAT TCT C 47°C<br />

Reverse<br />

TCT ATT ATT GCC TTA CTC ACA G<br />

D9S195<br />

Forward AGC TCA GCA CGG AGG G 53°C<br />

Reverse<br />

AGG GCA GGT TCC TAC AAA<br />

D9S258<br />

Forward GCT AGA GAT GCC CTTCGAG TG 53°C<br />

Reverse<br />

AGG ATT TAT AGA AAG TCC AAA ACC C<br />

D9S102<br />

Forward ATA GAC TTC CAG ACA GAT AG 50°C<br />

Reverse<br />

CCT CTC TCA TTC CTG GTA CT<br />

GSN<br />

Forward CAG CCA GCT TTG GAG ACA AC 50°C<br />

Reverse<br />

TCG CAA GCA TAT GAC TGT AA<br />

D9S1199<br />

Forward AAA AAT CAT GTG CAT CAA TTC C 53°C<br />

Reverse<br />

CCA GAG AAG CAG AAC CAA CG<br />

D9S1198<br />

Forward TGG GAG AGG GAA AAT GCT ATC 47°C<br />

Reverse<br />

GTA CTC CAG CCT GGG TGG<br />

2. Materials<br />

1. Xylene (BDH, UK).<br />

2. 100%, 95% and 70% alcohol (BDH, UK).<br />

3. 0.05% Toluidine blue dissolved in distilled water (Sigma Aldrich, UK).<br />

4. Leitz microscope and Leica micromanipulator (Leica, UK).<br />

5. Microdissecting needle and needle holder (Leica, UK).<br />

6. Protein digestion buffer (0.5% Tween in 50 mM Tris-EDTA, pH 8.5).<br />

7. 0.5 mg/mL proteinase K, add 450 µL of protein digestion buffer to 50 µL of 10 mg/m<br />

proteinase K dissolved in water (Sigma Aldrich, UK).<br />

8. 37°C water bath.<br />

9. MJ Research PTC-225 Peltier Thermal Cycler (Genetic Research Instrumentation, UK).<br />

10. 2 µM forward and reverse fluorescent-labeled primers per reaction (MWG-<strong>Bio</strong>tech, UK<br />

Ltd). Primers are present in the reaction at 0.2 µM. For individual primer sequences,<br />

refer to Table 1.


304 Edwards and <strong>Bartlett</strong><br />

11. Hot start Taq polymerase, 5 units/µL, reaction concentration is 0.5 units/reaction (Qiagen<br />

Ltd, UK).<br />

12. 10× reaction buffer and 25 mM MgCl 2 (Qiagen Ltd, UK); reaction concentration is<br />

1× reaction buffer and 3 mM MgCl 2 .<br />

13. dNTPs (10 mM; Advanced <strong>Bio</strong>tech); reaction concentration is 200 µm of each dNTP.<br />

14. PCR-grade water (Sigma Aldrich, UK).<br />

15. Formamide (BDH, UK).<br />

16. Loading buffer (50 mg/mL blue dextran, 25 mM EDTA).<br />

17. Genescan 400 HD size standard (rox), (ABI Perkin–Elmer, UK).<br />

18. Acrylamide (Sigma Aldrich, UK).<br />

19. Bisacrylamide (Sigma Aldrich, UK).<br />

20. Automated laser-activated fluorescent DNA sequencer, ABI 377 sequencer (Perkin<br />

Elmer, UK).<br />

21. Genescan and Genotyping software (Perkin Elmer, UK).<br />

3. Methods<br />

3.1. Microdissection of Archival DNA<br />

1. Archival formalin-fixed, paraffin-embedded TCCs are cut into 5-µm thick sections and<br />

put on to glass slides.<br />

2. Tissue sections are dewaxed by incubating in xylene (2 × 10 min).<br />

3. Tissue is rehydrated by incubating in a series of alcohol solutions, 100% alcohol<br />

(2 × 2 min), 95% alcohol (2 min), and 70% alcohol (2 min).<br />

4. Sections are stained in 0.05% Toluidine blue for 30 s.<br />

5. Areas of tumor cell are microdissected from 5-µm sections using a dissecting microscope<br />

and a micromanipulator.<br />

6. Dissected tissue is placed in an RNA/DNA free tube containing 12 µL of protein digestion<br />

buffer.<br />

7. DNA is extracted by addition of a 13 µL of protein digestion buffer containing proteinase<br />

K and incubating for 4 to 7 d at 37°C in a water (9).<br />

8. On removal from the water bath, proteinase K is inactivated by heating for 10 min at 95°C<br />

using a thermal cycler (see Note 1).<br />

9. Normal tissue is also dissected from the same section as the tumor, thus providing normal<br />

DNA to be used as a control in MIS and LOH analysis.<br />

10. LOH and MIS analysis is conducted using 1-µL aliquots of DNA without further<br />

purification.<br />

3.2. Primers and Loci Analyzed<br />

Primer sequences used for amplification of 9 microsatellites are found in Table 1.<br />

Two microsatellites (D9S126 and D9S259) are located on the short arm of chromosome<br />

9 spanning the 9p21 region, five microsatellites (D9S275, D9S195, D9S258, D9S103,<br />

and GSN) are located on the long arm of chromosome 9 spanning the 9q 32-33 region<br />

and two microsatellites (D9S1199 and D9S1198) are located on the long arm spanning<br />

the 9q 34 region. Primers were purchased from MWG-<strong>Bio</strong>tech UK Ltd, one primer<br />

from each pair was fluorescently labeled at the 5′ end.<br />

3.3. PCR<br />

The target sequences were amplified by PCR in 10-µL reactions, with each reaction<br />

containing the following.


Detection of MIS 305<br />

1. Archival DNA template (1 µL).<br />

2. 1× reaction buffer (1 µL of 10× reaction buffer); because the reaction buffer contained<br />

15 mM MgCl 2 , this gave a reaction concentration of 1.5 mM MgCl 2 .<br />

3. 2 µM reverse and forward primer (1 µL) to give a reaction concentration of 0.2 µM for<br />

each primer.<br />

4. 0.6 1 µL of 25 mM MgCl 2 to give a final reaction concentration of 3 mM MgCl 2 .<br />

5. 0.2 1 µL of 10 mM dNTP to give an reaction concentration of 200 µL each of dATP,<br />

dCTP, dGTP, and dTTP.<br />

6. 0.1 1 µL of 5 units/1 µL hot star Taq polymerase to give 0.5 units per reaction.<br />

7. Volume is made up to 10 1 µL using PCR-grade water.<br />

The reaction was started after a 15-min denaturation of DNA at 95°C. DNA<br />

amplification was performed in a thermal cycler as follows: 45 cycles of denaturation<br />

at 94°C for 30 s; annealing at 47, 50, or 53°C (Table 1) for 30 s and extension at 72°C<br />

for 1 min; followed by a final extension for 15 min at 72°C.<br />

The PCR is repeated using the same DNA template for each of the primer sets. On<br />

removal from the thermal cycler 2 µL of each of the 9 PCR products from each primer<br />

set are combined for each template and mixed thoroughly by vortexing.<br />

3.4. PCR Analysis<br />

1. PCR products are then prepared for gel electrophoresis.<br />

2. The following are combined in a 200-µL DNA/RNA free microfuge tube and vortexed:<br />

.5 µL of combined PCR products; 1 µL of formamide; 0.5 µL of loading buffer (50 mg/mL<br />

blue dextran, 25 mM EDTA); and 0.5 µL of commercial standard (Genescan 400 HD (rox),<br />

ABI Perkin–Elmer, Norwalk, CT).<br />

3. The mixture is then incubated at 96°C for 4 min and cooled on ice.<br />

4. The mixture (1.5 µL) is applied to a 4% arcrylamide/bisacrylamide gel and electrophoresed<br />

for 2 h on an automated laser-activated fluorescent DNA sequencer (Perkin–Elmer ABI<br />

377 sequencer).<br />

5. Fluorescent gel data were collected automatically during electrophoresis and analyzed<br />

using Genescan software. An example of the gel image is shown in Fig. 1. The data gained<br />

from Genescan is then exported into Genotyper and further analyzed.<br />

3.5. Assessment of Allele Loss and Microsatellite Instability<br />

Allele loss should be assessed as described by Dietmaier et al. (10). In heterozygous<br />

individuals, two alleles, that is, two PCR products of different size can be detected in<br />

normal DNA. Because PCR fragments of different sizes are amplified with different<br />

efficiencies, the ratio of allele peak heights is calculated in matched normal and tumor<br />

DNA. Peak heights of the longer length allele peaks are divided by the peak heights of<br />

the shorter length allele peaks (see Note 2). The ratio obtained in tumor DNA divided<br />

by the allele peak ratio of paired normal DNA gives a result range of 0.00 to 1.00 (see<br />

Note 3), that is, (tumor allele 1 peak height/tumor allele 2 peak height)/(normal allele<br />

1 peak height/normal allele 2 peak height).<br />

Theoretically a complete allele loss results is a value of 0 and both alleles retained<br />

results in a ratio of 1. In cases where the shorter length allele is lost the ratio obtained<br />

is greater than 1, this is therefore inverted (1/×) to obtain values within the 0.00 to<br />

1.00 range. A ratio below 0.65 represents an allele signal reduction of 35%, this is<br />

considered to be indicative of allele loss (see Note 4). This limit was chosen because


306 Edwards and <strong>Bartlett</strong><br />

Fig. 1. Example of a gel image.<br />

the tumor cell content after microdissection is assessed to be greater than 70% and inter<br />

assay variations of the detection system is below 5%. An example of LOH detection<br />

using the Geneotype ® Software can be seen in Fig. 2.<br />

A locus is described as unstable if a novel peak is present or if a peak has undergone<br />

a size shift after PCR amplification of tumor DNA compared with PCR amplification<br />

of paired normal DNA. A tumor is described as exhibiting MIS if at least 40% of the<br />

loci investigated are found to be unstable.<br />

4. Notes<br />

1. Inactivation of proteinase K is a very important step because failure to inactivate proteinase<br />

K will cause PCR to fail, with the proteinase K destroying the Taq polymerase.<br />

2. When analyzing Genotyper data, more than one peak may be present for each allele<br />

because of polymerase artifact stutter bands. It is therefore difficult to decide the size<br />

and the height of the peak to include in the calculation. In general, however, the sizes


Detection of MIS 307<br />

Fig. 2. The top graph shows LOH, as the peak at 104 base pairs is missing. However, the<br />

bottom graph demonstrates retention of both alleles.<br />

of the two alleles are assigned to the peaks of greatest height and the smaller peaks are<br />

interpreted as stutter bands.<br />

3. Tumor cells are not completely separated from adjacent normal cells by microdissection.<br />

Therefore, a major advantage of the fluorescence-based method is that the loss of alleles<br />

can be determined precisely by calculation of the ratio of the peak heights of normal<br />

and tumor alleles.<br />

4. Assays of samples with allele ratios in the border line range of 0.6 to 0.7 were repeated at<br />

least three times to ensure accuracy of results. Most repeated assays gave consistent results,<br />

and all <strong>info</strong>rmative markers gave similar allele ratios in the same sample.<br />

References<br />

1. King, B. L., Carcangiu, M.-L., Carter, D., Kiechle, M., Pfisterer, J., and Kaoinski, B. M.<br />

(1995) Microsatellite instability in ovarian neoplasms. Br. J. Cancer 72, 376–382.<br />

2. Sardi, I., Bartoletti, R., Occhini, I., Piazzini, M., Travaglini, F., Guazzelli, R., et al. (1999)<br />

Microsatellite alterations in superficial and locally advanced transitional cell carcinoma of<br />

the bladder. Oncol. Reports 6, 901–905.


308 Edwards and <strong>Bartlett</strong><br />

3. Sartor, O. and Sheng, A. (1997) Determinatin of CAG repeat length in the androgen receptor<br />

gene using frozen serum. Urology 49, 301–304.<br />

4. Bronner, C. E., Baker, A. M., Morrison, P. T., Warren, G., Smith, L. G., Lesloe, M. K., et al.<br />

(1994) Mutation in the DNA mismatch repair gene homologue hMLH1 is associated with<br />

hereditary non-polyposis colon cancer. Nature 368, 258–261.<br />

5. Rohrbach, H., Haas, J. C., Baretton, G. B., Hirschmann, A., Diebold, J., Behrendt, R. P.,<br />

et al. (1999) Microsatellite instability and loss of heterozygosity in prostatic carcinomas:<br />

Comparison of primary tumor and of corresponding recurrences after androgen deprivation<br />

therapy and lymph node metastases. Prostate 40, 20–27.<br />

6. Niederacher, D., Picard, F., van Roeyen, C. An, H. X., Bender, H. G., and Beckmann, W.<br />

(1997) Patterns of allelic loss on chromosome 17 in sporadic breast carcinomas dectected<br />

by fluorescent labelled microsatellite analysis. Genes Chromosomes Cancer 18, 181–192.<br />

7. Christensen, M., Sunde, L., Bolund, L., and Orntorft, T. F. (1999) Comparison of three<br />

methods of microsatellite detection. Scand. J. Clin. Lab. Invest. 59, 167–178.<br />

8. Serth, J., Kuczyk, M. A., Paeslack, U., Lichtinghagen, R., and Jonas, U. (2000) Quantitation<br />

of DNA extracted after microprepatation of cells from frozen and formalin-fixed tissue<br />

sections. Am. J. Pathol. 156, 1189–1196.<br />

9. Going, J. J. and Lamb, R. F. (1996) Practical histological microdissection for PCR analysis.<br />

J. Pathol. 79, 121–124.<br />

10. Dietmaier, W., Riedlinger, W., Kohler, A., Wegele, P., Beyser, K., Sagner, G., et al. (1999)<br />

Detection of microsatellite instability (MSI) and loss of heterozygosity (LOH) in colorectal<br />

tumours by fluorescence based multiplex microsatellite PCR. <strong>Bio</strong>chemica 2, 43– 46.


Shadow Band Synthesis 309<br />

44<br />

Reduction of Shadow Band Synthesis During<br />

PCR Amplification of Repetitive Sequences<br />

from Modern and Ancient DNA<br />

Wera M. Schmerer<br />

1. Introduction<br />

Repetitive sequences like short tandem repeat (STR) loci are generally referred to<br />

as slippery DNA (1). They owe this nickname to a characteristic leading to slippage<br />

within the primer-template complex during PCR elongation of the new strand (2,3),<br />

resulting in the synthesis of byproducts shortened by one repeat unit compared with<br />

the original sequence. The generation of these so-called shadow bands (4) is a wellknown<br />

problem connected with the amplification of repetitive DNA, complicating the<br />

genotype analysis of modern (e.g., ref. 5), forensic (6), and ancient (7,8) specimens.<br />

In some applications, the occurrence of this artifact makes it necessary to develop<br />

guidelines for allele designation (6,9).<br />

The intensity of these byproducts increases with the degradation of the target DNA<br />

(8,10), and therefore represents a particular problem concerning the analysis of highly<br />

degraded DNA as in genotyping of forensic (e.g., refs. 11,12) and ancient DNA<br />

(7,13,14). In amplification products of highly degraded or ancient DNA, the intensity<br />

of a shadow band can exceed the peak height or band intensity of the original allele.<br />

As artifact alleles (7,15) they can lead to mistyping of amplification products (cf. also<br />

refs. 8,16). Because an amplification product is not necessarily affected in each case,<br />

this phenomenon can even result in seemingly different genotypes for independent<br />

amplification products of the same sample (7,15,16).<br />

To improve the reproducibility of amplification results and consequently decrease<br />

the probability of mistyping by reducing the generation of this artifact, the optimization<br />

of the PCR amplification process itself represents the most important strategy besides<br />

optimization of the extraction of the DNA used as template (17–19).<br />

The findings presented in the following resulted from a study within the scope<br />

of which different strategies to optimize PCR amplification of repetitive DNA were<br />

investigated with reference to their effect on the generation of shadow bands (cf.<br />

ref. 19 Schmerer, manuscript in preparation). The model locus investigated was<br />

HUMVWA31/A (20), the amplifications of which show a high tendency to accumulate<br />

shadow bands compared with other STR loci (11,12,21). Amplifications were performed<br />

on DNA Thermal Cycler (TC1, Perkin–Elmer Cetus). For a detailed presentation of<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

309


310 Schmerer<br />

the investigation concerned, please refer to Schmerer (19) and Schmerer (manuscript<br />

in preparation).<br />

2. Materials<br />

1. InViTAQ ® DNA polymerase (Invitek).<br />

2. NH 4 reaction buffer (Invitek). 10× buffer: 160 mM (NH 4 ) 2 SO 4 , 500 mM Tris-HCl<br />

(pH 8.3 at 25°C), Tween ® 20.<br />

3. Betaine (3 M solution with sterile water also used for set up of the reaction mix).<br />

4. Betaine (2 M) + 10% dimethyl sulfoxide (DMSO; solution with sterile water).<br />

5. Bovine serum albumin (BSA; 125 µg/mL solution with sterile water).<br />

6. dNTP-mix composed according to the sequence amplified (e.g., for HUMVWA31/A with<br />

a A/TG/C ratio of 1.91, a stock of 121 µM each of dGTP, dCTP, and 220 µM each of<br />

dATP and dTTP were used).<br />

3. Methods<br />

Each of the following variations of a standard PCR amplification protocol resulted<br />

in reduced accumulation of shadow bands. They might be applied either soldiery or<br />

in any combination.<br />

3.1. Denaturation Time<br />

A reduction of denaturation time to 15 to 30 s results in a 28% decrease in shadow<br />

bands compared with the standard denaturation time of 1 min by decreasing the<br />

occurrence of additional degradation of the template DNA (cf. ref. 22) because of the<br />

shorter incubation period at the high temperature of 94°C.<br />

3.2. Elongation Step<br />

Applying an elongation time of 1 min, which is the standard used for synthesis<br />

of a PCR product up to 1 kb, the lowest intensity of shadow bands was found with<br />

an elongation temperature of 68°C, resulting in a 20.4% reduction compared with<br />

the generally applied temperature of 72°C. This reduction could be even increased to<br />

24.7% by doubling the time for synthesis. The highest reduction of the generation of<br />

this artefact (–30.9%) could be achieved with an elongation at 70°C for 2 min.<br />

3.3. Polymerase and Composition of Reaction Buffer<br />

Comparing different polymerases and polymerase mixes respective, the lowest<br />

intensity of shadow bands was observed with InViTAQ ® (Invitek) and the CombiPoil ®<br />

polymerase mix (Invitek) in combination with OptiPerform Buffer (Invitek) with<br />

a reduction of the artefact by up to 24 and 21.4% compared with AmpliTaq Gold<br />

(Perkin–Elmer) in combination with GeneAmp ® buffer (Perkin–Elmer). Applying<br />

InViTaq, a replacement of the generally used KCl-reaction buffer, by an optimized<br />

NH 4 buffer resulted in a 22% reduction in artefact accumulation applying the same<br />

polymerase. Further differences of the two buffer systems consist in a slightly elevated<br />

pH value of 8.8 (+0.5) and an addition of 0.01% Tween-20 in case of the NH 4 buffer.<br />

The use of a polymerase displaying a 3′→5′ exonuclease proof-reading ability<br />

showed no positive effects concerning the reduction of shadow band accumulation,<br />

neither applied alone, nor in combination with a Taq polymerase.


Shadow Band Synthesis 311<br />

3.4. A/TG/C Ratio of the dNTP-Mix<br />

Changing the composition of the dNTP-mix from the usual equimolar concentration<br />

of nucleotides to a A/TG/C ratio of 1.91 corresponding to the composition of<br />

the general sequence of HUMVWA31/A, the locus amplified, resulted in a decrease<br />

in shadow band generation by 7.3%. Changes in amplification efficiency were not<br />

observed, neither in processing modern nor ancient DNA. Equability of the amplification<br />

of both alleles belonging to a heterozygous genotype was slightly improved.<br />

3.5. Betaine (N,N,N-trimethyl glycine)<br />

The presence of betaine in a concentration of 0.5 to 2 M reduced the accumulation<br />

of shadow bands, with a maximum reduction of 15.5% at 0.5 M. Concentrations lower<br />

than 0.5 showed an increase in artifact production. Concerning the amplification of<br />

ancient DNA, low concentrations of betaine (0.25–1 M) resulted in an increase in<br />

product yield because of the neutralizing effect of this reagent against inhibitory<br />

substances (cf. ref. 23) frequently present within ancient DNA extracts. Beyond 1.5 M,<br />

an addition of the inhibitory effect of betaine at elevated concentrations (24) and the<br />

inhibition caused by co-extracted impurities occurred, resulting in partial inhibition<br />

of amplification.<br />

3.6. Betaine Combined with DMSO<br />

Like betaine, low concentrations of both betaine and DMSO resulted in a slight<br />

increase in intensity of shadow bands. A reduction of the artifact was observed at<br />

concentrations of 0.4 to 1 M betaine combined with 2 to 5% DMSO, with a maximum<br />

decrease of shadow band intensity by 12.6% at 0.8 M betaine and 4% DMSO. In<br />

ancient DNA amplifications, the presence of these reagents increased product yield and<br />

reproducibility between multiple amplifications of the same sample (see Note 1).<br />

3.7. BSA<br />

The addition of BSA in a concentration of 10 to 25 µg/mL resulted in reduced<br />

intensity of artifact bands with a maximum at 25 µg/mL by 21.2% accompanied by an<br />

increased equability of amplification of a heterozygote genotype. In addition to this,<br />

the presence of BSA showed to increase the efficiency of ancient DNA amplifications<br />

by neutralizing inhibitory substances (25) that consequently resulted in higher yield<br />

of specific product. A concentration of 25 to 50µg/mL was determined as optimal for<br />

the amplification of ancient DNA.<br />

4. Notes<br />

1. Also investigated concerning their impact in the generation of shadow bands were further<br />

reagents commonly used as PCR-enhancing additives, such as DMSO, glycerol, and<br />

formamide. In amplifications of the locus concerned the addition of DMSO, as well as<br />

formamide in different concentrations resulted in an increase in shadow band accumulation.<br />

Glycerol, however, did not show any effect, neither on modern, nor ancient DNA<br />

amplification.


312 Schmerer<br />

References<br />

1. Kunkel, T. A. (1993) Nucleotide repeats. Slippery DNA and diseases. Nature 365,<br />

207–208.<br />

2. Levison, G. and Gutman, G. A. (1987) Slipped-strand mispairing: a mayor mechanism for<br />

DNA sequence evolution. Mol. <strong>Bio</strong>l. Evol. 4, 203–221.<br />

3. Schlötterer, C. and Tautz, D. (1992) Slippage synthesis of simple sequence DNA. Nucleic<br />

Acids Res. 20, 211–215.<br />

4. Weber, J. L. and May, P. E. (1989) Abundant class of human polymorphisms which can be<br />

typed using the polymerase chain reaction. Am. J. Hum. Genet. 44, 388–396.<br />

5. Bluteau, O., Legoix, P., Laurent-Puig, P., and Zucman-Rossi, J. (1999) PCR-based genotyping<br />

can generate artifacts in LOH analyses. <strong>Bio</strong>Techniques 27, 1100–1102.<br />

6. Gill, P., Sparkes, R., and Kimpton, C. (1997) Development of guidelines to designate alleles<br />

using a STR multiplex system. Forensic Sci. Int. 89, 185–197.<br />

7. Schmerer, W. M., Hummel, S., and Herrmann, B. (1997) Reproduzierbarkeit von aDNAtyping.<br />

Anthrop. Anz. 55, 199–206.<br />

8. Ivanov, P. L. and Isaenko, M. V. (1999) Identification of human decomposed remains<br />

using the STR systems: effect on typing results, in Proceedings of the second European<br />

symposium on human identification 1998. Promega Corporation, Innsbruck, Austria.<br />

9. Fourney, R. M., Fregau, C. J., Bowen, J. H., Bowen, K. L., Shutler, G. G., Elliot, J. C., et al.<br />

(1995) Interpretation guidelines for fluorescent automated detection of STRs: Defining the<br />

allele and the limits of detection, in Proceedings from the Sixth International Symposium on<br />

Human Identification 1995. Promega Corporation, Innsbruck, Austria.<br />

10. Murray, V., Monchawin, C., and England, P. R. (1993) The determination of the sequences<br />

present in the shadow bands of a dinucleotide repeat PCR. Nucleic Acids Res. 21,<br />

2395–2398.<br />

11. Kimpton, C., Fisher, D., Watson, S., Adams, M., Urquhard, A., Lygo, J., et al. (1994)<br />

Evaluation of an automated DNA profiling system employing multiplex amplification of<br />

four tetrameric STR loci. Int. J. Leg. Med. 106, 302–311.<br />

12. Lygo, J. E., <strong>John</strong>son, P. E., Holaway, D. J., Woodroffe, S., Whitaker, J. P., Clayton, T. M.,<br />

et al. (1994) The validation of short tandem repeat (STR) loci for use in forensic casework.<br />

Int. J. Leg. Med. 107, 77–89.<br />

13. Ramos, M. D., Lalueza, C., Girbau, E., Perez-Perez, A., Quevedo, S., Turbon, D., et al.<br />

(1995) Amplifying dinucleotide microsatellite loci from bone and tooth samples of up to<br />

5000 years of age: More inconsistency than usefulness. Hum. Genet. 96, 205–212.<br />

14. Schultes, T., Hummel, S., and Herrmann, B. (1997) Recognizing and overcoming inconsistencies<br />

in microsatellite typing of ancient DNA samples. Ancient <strong>Bio</strong>mol. 1, 227–233.<br />

15. Schmerer, W. M. (1996) Reproduzierbarkeit von Mikrosatelliten-DNA-Amplifikation und<br />

Alleldetermination aus bodengelagertem Skelettmaterial. Unpublished diploma-thesis,<br />

Göttingen.<br />

16. Taberlet, P., Griffin, S., Goossens, B., Questiau, S., Manceau, V., Escaravage, N., et al.<br />

(1996) Reliable genotyping of samples with very low DNA quantities using PCR. Nucleic<br />

Acids Res. 24, 3189–3194.<br />

17. Schmerer, W. M., Hummel, S., and Herrmann, B. (1999) Optimized DNA extraction to<br />

improve reproducibility of short tandem repeat genotyping with highly degraded DNA as<br />

target. Electrophoresis 20, 1712–1716.<br />

18. Schmerer, W. M., Hummel, S., and Herrmann, B. (2000) STR-genotyping of archaeological<br />

human bone: Experimental design to improve reproducibility by optimisation of DNA<br />

extraction. Anthrop. Anz. 58, 29–35.<br />

19. Schmerer, W. M. (2000) Optimierung der STR-Genotypenanalyse an Extrakten alter DNA<br />

aus bodengelagertem menschlichen Skelettmaterial. Cuvillier Verlag, Göttingen.


Shadow Band Synthesis 313<br />

20. Kimpton, C. P., Walton, A., and Gill, P. (1992) A further tetranucleotide repeat polymorphism<br />

in the vWF gene. Hum. Mol. Genet. 1, 287.<br />

21. Bacher, J. and Schumm, J. W. (1998) Development of highly polymorphic pentanucleotide<br />

tandem repeat loci with low stutter. Profiles DNA 2, 3–6.<br />

22. Cheng, S., Fockler, C., Barnes, W. M., and Higuchi, R. (1994) Effective amplification of<br />

long targets from cloned inserts and human genomic DNA. Proc. Natl. Acad. Sci. USA<br />

91, 5695–5699.<br />

23. Baskaran, N., Kandpal, R. P., Bhargava, A. K., Glynn, M. W., Bale, A., and Weissman, S. M.<br />

(1996) Uniform amplification of a mixture of deoxyribonucleic acids with varying GC<br />

content. Genome Res. 6, 633–638.<br />

24. Weissensteiner, T. and Lanchbury, J. S. (1996) Strategy for controlling preferentiual<br />

amplification and avoiding false negatives in PCR Typing. <strong>Bio</strong>techniques 21, 1102–1108.<br />

25. Pääbo, S., Gifford, J. A., and Wilson, A. C. (1988) Mitochondrial DNA sequences from a<br />

7000-year old brain. Nucleic Acids Res. 20, 9775–9787.


314 Schmerer


DOP-PCR 315<br />

45<br />

Degenerate Oligonucleotide-Primed PCR<br />

Michaela Aubele and Jan Smida<br />

1. Introduction<br />

The amount of genomic DNA available for genetic studies can often be limiting.<br />

Degenerated oligonucleotide-primed polymerase chain reaction (DOP-PCR) is an<br />

appropriate method for overcoming these limitations by efficiently performing whole<br />

genome amplification. The DOP-PCR technique is increasingly being applied for<br />

simultaneous amplification of multiple loci in target DNA using oligonucleotide<br />

primers of partially degenerate sequences (1). Contrary to other PCR-based general<br />

amplification methods (Alu-PCR or IRS-PCR), DOP-PCR is a species-independent<br />

technique for the amplification of small amounts of DNA.<br />

Briefly, the use of one single degenerated primer allows random initial priming all<br />

over the target DNA during the first five cycles of PCR at low annealing temperatures.<br />

In subsequent cycles, more stringent conditions are then used to amplify the first PCR<br />

products. DOP-PCR technique is recently being applied to produce sufficient amounts<br />

of DNA from clinical tumor samples, in forensic analyses, prenatal diagnosis, and<br />

many other investigations. Slightly different protocols have been used and optimized<br />

worldwide. The protocol described here is adapted for use with microdissected,<br />

formalin-fixed, paraffin-embedded human tissue in comparative genomic hybridization<br />

analysis. A comparison of several DOP-PCR variations is also given by Larsen<br />

et al. (2).<br />

2. Materials<br />

2.1. Deparaffination and Sampling of Material<br />

1. Xylene (100%).<br />

2. Ethanol (100%, 70%, 50%).<br />

3. Distilled water.<br />

4. Mayer’s Hemalaun solution (Merck 109 249, Darmstadt, Germany).<br />

5. Laser buffer: 100 mM Tris-HCl, pH 7.5.<br />

6. Proteinase K: 10 mg/mL proteinase K, pH 7.5 (Merck, Darmstadt, Germany).<br />

7. Qiagen DNA Mini Kit (Qiagen, Hilden, Germany).<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

315


316 Aubele and Smida<br />

2.2. Chemicals Required for DOP-PCR<br />

1. DOP-primer: 5′- CCG ACT CGA GNN NNN NAT GTG G - 3′, with N = A, C, G, or T in<br />

approx equal proportions, (<strong>Bio</strong>metra, Göttingen, Germany).<br />

2. Topoisomerase I (Life Technologies, Eggenstein, Germany).<br />

3. Taq DNA Polymerase (Perkin–Elmer Life Sciences, Maryland).<br />

4. DNTPs (10 mM each of dATP, dCTP, dGTP, and TTP; Sigma-Aldrich, Steinheim,<br />

Germany).<br />

5. 10× amplification buffer: 500 mM KCl, 200 mM Tris-HCl, pH 8.4.<br />

6. 50 mM MgCl 2 .<br />

3. Method<br />

3.1. Preparation of Tumor Samples<br />

1. Dewax 5- to 10-µm thick paraffin sections for 30 min in a Coplin jar with Xylene.<br />

2. Transfer slides to a coplin jar with fresh xylene for 5 min.<br />

3. Transfer slides to a Coplin jar with 100% EtOH for 5 min.<br />

4. Transfer slides to a coplin jar with 70% EtOH for 5 min.<br />

5. Transfer slides to a coplin jar with 50% EtOH for 5 min.<br />

6. Transfer slides to a coplin jar with distilled water for 5 min. For sampling of small lesions<br />

or even single cells by microdissection, see chapter 2.1.3<br />

3.2. DOP-PCR Procedure (see Note 1)<br />

An optional Topoisomerase step may be performed (see Note 2).<br />

1. Prepare PCR mix (50-µL volume, see Note 3):<br />

Template<br />

Pretreated cell sample in 20 µL of laser<br />

buffer or up to 100 ng isolated DNA<br />

Primer 6-MW (0.1 nmol/µL, see Note 4) 1.0 µL<br />

10× amplification buffer 5.0 µL<br />

MgCl 2 3.4 µL<br />

dNTP (40 mM) 1.2 µL<br />

Taq Polymerase (5 U/µL) 0.8 µL<br />

H 2 O Fill up to final volume of 50 µL<br />

2. Set up cycler conditions: Start PCR in a thermal cycler in a 50-µL reaction volume using<br />

the following conditions (see Note 5).<br />

Denaturation 10 min at 94°C<br />

Followed by five cycles: 11 min at 94°C<br />

(Low stringency) 11.3 min at 30°C<br />

13 min transition 30–72°C<br />

13 min extension at 72°C<br />

Followed by 35 cycles: 11 min at 94°C<br />

(High stringency) 11 min at 62°C<br />

13 min at 72°C<br />

Final extension 10 min at 72°C<br />

3. Store PCR product at –20°C or continue to next step.<br />

4. Use a 5-µL aliquot of the PCR product to check the DNA fragment size on a 3% agarose<br />

ethidium bromide gel (see also Fig. 1),<br />

5. Determine DNA amounts by fluorometric or photometric measurements.


DOP-PCR 317<br />

Fig. 1. DOP-PCR products reveal a typical smear ranging from about 100 to 2500 bp. Lane 1:<br />

marker pBR 322, HaeIII; lane 2: positive control (PCR product with template DNA and a<br />

gene-specific primer (β-Actin); lane 3: negative control (PCR product without template and<br />

with degenerate primer); lane 4: empty; lanes 5 to 9: DNA samples from formalin-fixed,<br />

paraffin-embedded tissue that were amplified by DOP-PCR.<br />

4. Notes<br />

1. As in all preparations that require handling of small cell and/or small DNA amounts, the<br />

PCR mix should be set up in a laminar flow to avoid any contamination. Furthermore,<br />

because of the species-independent primer, a strict contamination-free working is<br />

prerequisite.<br />

2. Tumor samples may be pretreated for 30 min with 2 units Topoisomerase I at 37°C to relax<br />

the template DNAs. Stop activation of topoisomerase at 90°C for 10 min and chill on<br />

ice, then add Taq DNA Polymerase and start PCR. This additional step is recommended<br />

for isolated fresh material but is not necessary for degraded DNA, as is the case for<br />

formalin-fixed, paraffin-embedded tissue.<br />

3. All PCRs should be controlled for possible contamination by using at least one negative<br />

control (vial without template DNA), and one positive control (vial with template DNA<br />

and a gene specific primer, e.g., β-actin; Fig. 1).<br />

4. The most commonly used DOP-PCR primer is 6-MW, originally described by Telenius (1),<br />

a degenerate primer with 6 specific 3′ bases and a XhoI site at its 5′ end (the same primer<br />

has been described as UW4B, UN-1, or MTE1 by different authors).<br />

5. More than 40 cycles in DOP-PCR should be avoided. Based on the authors’ experiences,<br />

artificial results could arise. For this reason, labeling of DNA should be performed, for<br />

example, by Nick translation instead of a further DOP-PCR.


318 Aubele and Smida<br />

References<br />

1. Telenius, H., Carter, N. C., Bebb, C. E., Nordenskjöld, M., Ponder, B. A. J, and Tunnacliffe,<br />

A. (1992) Degenerate oligonucleotide-primed PCR: General amplification of target DNA<br />

by a single degenerate primer. Genomics 13, 718–725.<br />

2. Larsen, J., Ottesen, A. M., Lundsteen, C., Leffers, H., and Larsen, J. K. (2001) Optimization<br />

of DOP-PCR amplification of DNA for high resolution comparative genomic hybridization<br />

analysis. Cytometry 44, 317–325.


Mutation Detection Using RT-PCR-RFLP 319<br />

46<br />

Mutation Detection Using RT-PCR-RFLP<br />

Hitoshi Nakashima, Mitsuteru Akahoshi, and Yosuke Tanaka<br />

1. Introduction<br />

Genetic analysis by restriction fragment length polymorphism (RFLP) is one of<br />

the most common methods used to examine nucleic acids for the presence of known<br />

sequence variants. A segment that is to be searched for a mutation is amplified from<br />

genomic DNA or cDNA, digested by the appropriate restriction enzyme, and then<br />

separated by agarose gel electrophoresis. Although RFLP analysis is a highly sensitive<br />

method that is easy to apply for the screening of known sequence variants, many<br />

common polymorphisms are the result of single-base substitutions that fail to create<br />

or remove any restriction site and, therefore, these cannot be readily typed by simple<br />

polymerase chain reaction (PCR) and RFLP analysis. However, the use of a mismatch<br />

PCR primer to artificially create a restriction site in the amplified product make it<br />

possible to overcome this disadvantage (1,2). The mismatch primer contains a oneor<br />

two-base mismatch near its 3′ end such that the amplified product incorporates or<br />

removes a restriction site for the appropriate endonuclease in the presence of a base<br />

substitution (3–5). The protocol for this method is the same as standard PCR.<br />

2. Materials<br />

1. RNA extracted from cells or tissues of interest.<br />

2. Primers: stock solutions are at 100 µM in H 2 O; working solutions are at 10 µM in H 2 O.<br />

The following primer pairs yield a 164-bp amplicon. The sites of mismatch are in capital<br />

letters (3). Forward mismatch primer: 5′-ctc cta ccc ctt gtc atg cag gAt-3′; reverse primer:<br />

5′-gtt aaa aca ggg acc tgt ggc atg-3′.<br />

3. 10× PCR buffer: 500 mM KCl, 100 mM Tris-HCl, pH 8.3.<br />

4. MgCl 2 (25 mM) in distilled water.<br />

5. dNTPs (10 mM each of dATP, dGTP, dCTP, and TTP).<br />

6. RNase inhibitor (20 U/µL).<br />

7. Random hexamers (50 µM).<br />

8. MuLV reverse transcriptase (50 U/µL).<br />

9. AmpliTaq DNA polymerase (5 U/µL; Roche Molecular Systems, Inc. Branchburg, NJ).<br />

10. Restriction enzyme FokI (10 U/µL) and digestion buffer.<br />

11. DEPC-treated sterile H 2 O.<br />

12. Sterile mineral oil.<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

319


320 Nakashima, Akahoshi, and Tanaka<br />

13. 80% ethanol.<br />

14. PEG-NaCl solution (20% PEG6000, 2.5 M NaCl).<br />

15. Microcentrifuge tube (0.5 mL).<br />

16. Microcentrifuge tube (1.5 mL).<br />

17. Equipment and reagents for 4% agarose gel electrophoresis.<br />

18. TBE: 89 mM Tris, 89 mM boric acid, and 2 mM EDTA (pH 8.3).<br />

19. Ethidium bromide (10 µg/mL).<br />

20. Loading buffer: 0.4% bromophenol blue, 0.4% xylene cyanol FF, 50% glycerol in H 2 O.<br />

3. Method<br />

To provide a DNA target for the PCR, cDNA is synthesized using reverse transcriptase.<br />

1. Place 1 µg of RNA in a 0.5-mL microcentrifuge tube, and add 4 µL of 25 mM MgCl 2 ,<br />

2 µL of 10× PCR buffer, 8 µL of 10 mM dNTP, 1 µL of RNase inhibitor, 1 µL of random<br />

hexamers, 1 µL of MuLV reverse transcriptase, and DEPC-treated sterile H 2 O to form<br />

final total volume of 20 µL.<br />

2. Mix and stand for 10 min at room temperature. Add two drops of mineral oil. Briefly<br />

spin down the reaction mixture.<br />

3. Incubate the reaction mixture for 15 min at 42°C, then for 5 min at 99°C, and finally<br />

for 5 min at 4°C.<br />

4. Add 4 µL of 25 mM MgCl 2 , 8 µL of 10× PCR buffer, 65.5 µL of sterile H 2 O, 0.5 µL of<br />

AmpliTaq DNA polymerase, and 1 µL of forward primer (10 µM) and reverse primer<br />

(10 µM ) to the RT product and mix. Briefly spin down the reaction mix.<br />

5. Perform the PCR by subjecting the reaction mixture to an initial denaturation of 3 min<br />

at 94°C, 35 cycles of 1 min at 94°C and 1 min at 60°C, followed by a final extension<br />

of 7 min at 72°C.<br />

6. Discard the mineral oil and transfer the PCR product to a 1.5-mL microcentrifuge tube.<br />

7. Add 60 µL of PEG-NaCl solution to the PCR product. Mix and stand for 10 min at 37°C.<br />

8. Centrifuge at 12,000g for 10 min at 4°C.<br />

9. The DNA precipitate should be visible as a pellet at the bottom of the tube. Remove the<br />

supernatant and wash the pellet with 1 mL of 80% ethanol.<br />

10. Centrifuge at 7500g for 5 min at 4°C.<br />

11. Carefully remove all the supernatant and air dry the pellet.<br />

12. Dissolve the pellet in 100 µL of H 2 O.<br />

3.1. Restriction Enzyme Digestion<br />

1. Mix 5 µL of PCR product, 1 µL of 10× reaction buffer, 1 µL of the appropriate restriction<br />

enzyme (1–10 U), and 3 µL of H 2 O in a 1.5-mL microcentrifuge tube.<br />

2. Incubate the reaction mixture for more than 2 h at the optimum temperature for enzyme<br />

activity.<br />

3.2. Analysis of RFLP<br />

1. Prepare an appropriate concentration of agarose gel containing 0.2 µg/mL of ethidium<br />

bromide for electrophoresis with TBE buffer.<br />

2. Load the gel with 10 µL of digested PCR product mixed with 1 µL of 10× loading buffer,<br />

5 µL of undigested PCR product mixed with 4 µL of H 2 O, and 1 µL of 10× loading buffer<br />

and a DNA size standard.<br />

3. Perform electrophoresis and inspect the gel under an ultraviolet transilluminator (Fig. 1).


Mutation Detection Using RT-PCR-RFLP 321<br />

Fig. 1. G88A within IFNGR1 is detected by FokI digestion of the PCR product using a 1-bp<br />

mismatch primer. The upstream primer contains a 1-bp mismatch just proximal to its 3′ end<br />

such that the 164-bp amplified product incorporates a restriction site for FokI in the presence of<br />

guanine at nucleotide 88 (a allele) but not in the presence of adenine (b allele). A PCR product<br />

of genotype ab is cleaved by endonuclease FokI and shows three bands, 164, 130, and 34 bp, on<br />

the ethidium bromide-stained 4% agarose gel.<br />

Lowercase nucleotides represent the mismatch base. Underlined nucleotides represent the<br />

polymorphism site. Line 1: Part of the registered IFNGR1 cDNA sequence; line 2: the upstream<br />

primer for mismatch PCR; line 3: restriction enzyme Fok I recognition site.<br />

4. Notes<br />

We were searching for amino acid polymorphisms in cytokine receptors. At first<br />

we used the RT-PCR SSCP method for cDNA sequence. After confirmation of the<br />

base substitutions of SSCP-positive samples, we designed primers for population<br />

screening by RFLP analysis. Although RT-PCR RFLP is an easy and sensitive method<br />

for examining whether an already-known base substitution is present within the sample<br />

cDNA, there are some disadvantages involved. Concerning RT-PCR, the various kinds<br />

of amplicons resulting from alternative splicing could prove to be an obstacle for<br />

analysis. If the genomic DNA sequence is available, then analysis using PCR products<br />

templated with genomic DNA is recommended. Changing the primer setting position<br />

would be another solution. Concerning RFLP, this analysis is based on the fact that<br />

each restriction enzyme recognizes a specific DNA sequence, and incomplete digestion<br />

leads to a false-negative result. The causes of incomplete digestion are low enzyme<br />

activity (in cases with long-term usage of the restriction enzyme), an extremely<br />

small quantity of the enzyme compared with that of the DNA sample, reaction under<br />

the wrong conditions, and the use of the wrong reaction buffer. Performance of the


322 Nakashima, Akahoshi, and Tanaka<br />

same reactions using positive and negative control templates in parallel with sample<br />

examinations is recommended in order to check for incomplete digestion.<br />

References<br />

1. Russ, A. P., Maerz, W., Ruzicka, V., Stein, U., and Gross, W. (1993) Rapid detection of<br />

the hypertension-associated Met235-Thr allele of the human angiotensinogen gene. Hum.<br />

Mol. Genet. 2, 609–610.<br />

2. Hingorani, A. R. and Brown, J. M. (1995) A simple molecular assay for the C1166 variant of<br />

the angiotensin II type 1 receptor gene. <strong>Bio</strong>chem. <strong>Bio</strong>phys. Res. Commun. 213, 725–729.<br />

3. Tanaka, Y., Nakashima, H., Hisano, C., Kohsaka, T., Nemoto, Y., Niiro, H., et al. (1999)<br />

Association of the interferon-g receptor variant (Val14Met) with systemic lupus erythematosus.<br />

Immunogenetics 49, 266–271.<br />

4. Shibata, S., Asano, Y., Yokoyama, T., Shimoda, K., Nakashima, H., Okamura, S., et al.<br />

(1998) Analysis of the granulocyte colony-stimulating factor receptor gene structure using<br />

PCR-SSCP in myeloid leukemia and myelodysplastic syndrome. Eur. J. Haematol. 60,<br />

197–201.<br />

5. Iwata, I., Nagafuchi, S., Nakashima, H., Kondo, S., Koga, T., Yokogawa, Y., et al. (1999)<br />

Association of polymorphism in the NeuroD/Beta2 gene with type 1 diabetes in the<br />

Japanese. Diabetes 48, 416– 419.


ARMS PCR 323<br />

47<br />

Multiplex Amplification Refractory<br />

Mutation System for the Detection<br />

of Prothrombotic Polymorphisms<br />

David Stirling<br />

1. Introduction<br />

First described by Newton and colleagues in 1989 (1), amplification refractory mutations<br />

system (ARMS) has become a standard technique that allows the discrimination<br />

of alleles that differ by as little as 1 bp. The system is simple, reliable, and nonisotopic.<br />

It clearly distinguishes heterozygotes at a locus from homozygotes for either allele. The<br />

system requires neither restriction enzyme digestion, nor allele-specific oligonucleotides<br />

as conventionally applied, nor the sequence analysis of polymerase chain reaction<br />

(PCR) products. The basis of the system is that oligonucleotides with a mismatched<br />

3′-residue will not function as primers in the PCR under appropriate conditions.<br />

A standard ARMS PCR consists of two complementary reactions (two tubes) and<br />

uses 3 primers. One primer is constant and complementary to the template in both reactions,<br />

and the other primers differ at their 3′ terminal residues and are specific to either<br />

the wild-type DNA sequence or the mutated sequence at a given base—only one of these<br />

primers is used per tube. If the sample is homozygous mutant or homozygous wildtype<br />

amplification will only occur in one of the tubes, if the sample is heterozygous<br />

amplification will be seen in both tubes.<br />

Here, we report our protocol for the multiplex ARMS detection of the polymorphisms<br />

in clotting factor V and II associated with increased risk of thrombosis (2,3). A portion<br />

of the factor IX gene is amplified as an internal positive control.<br />

2. Materials<br />

1. Thermal cycler.<br />

2. Plate mixer.<br />

3. Vertical PAGE system.<br />

4. Electrophoresis Power Pack.<br />

5. Thermowell 96-well plate and covers.<br />

6. Round tips (200 µL).<br />

7. Flat-cap PCR tubes (0.5 mL).<br />

8. Thermostable DNA Taq polymerase.<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

323


324 Stirling<br />

9. TAE (20×): 484 g of Tris, 114 mL of glacial acetic acid, 20 mL of 0.5 M EDTA. Dissolve<br />

in 3 L of distilled water then make up to 5 L with distilled water.<br />

10. TAE (1×): 120 dilution of 20× TAE in distilled water.<br />

11. 8% polyacrylamide gel.<br />

12. dNTPs: supplied separately as four different types at concentrations of 100 mM. For use<br />

at 5 mM, take 100 µL of each stock dNTP and add to one tube 7600 µL of sterile distilled<br />

water. Aliquot and store at –20°C.<br />

Oligo F2 Wild<br />

5′-CAC TGG GAG CAT TGA GGA TC-3′<br />

Oligo F2 Mutant<br />

5′-CAC TGG GAG CAT TGA GGA TT-3′<br />

Oligo F2 Consensus 5′-TCT AGA AAC AGT TGC CTG GC-3′<br />

Oligo F5 Wild<br />

5′-CAG ATC CCT GGA CAG ACG-3′<br />

Oligo F5 Mutant<br />

5′-CAG ATC CCT GGA CAG ACA-3′<br />

Oligo F5 Consensus 5′-TGT TAT CAC ACT GGT GCT TAA-3′<br />

Oligo F9 Forward<br />

5′-CTC CTG CAG CAT TGA GGG AGA TGG ACA TT-3′<br />

Oligo F9 Reverse<br />

5′-CTC GAA TTC GGC AAG CAT ACT CAA TGT AT-3′<br />

13. Oligonucleotides. The oligonucleotides must be diluted to 100 pmol/µL on arrival. Aliquot<br />

in 20-µL volumes and store at –20°C.<br />

3. Methods<br />

1. Using sterile pipet tips, take 1.0 µL of each sample into identified wells in plate, one<br />

labeled ‘W’ (wild type), the other ‘M’ (mutant). Ensure that the DNA sample is pipetted<br />

directly into the bottom of the respective well.<br />

2. Prepare PCR mastermixes W and M with the following: DNTP (n × 1.5 µL); 10× polymerase<br />

reaction buffer (n × 2.5 µL); oligonucleotides (each (n × 0.5 µL); MgCl 2 (25 mM;<br />

(n × 2 µL); Taq polymerase (n × 0.2 µL); Sterile distilled water (n × 15.4 µL); where<br />

n = number of samples/controls + 2.<br />

3. Mix mastermixes W and M and add 24 µL to each corresponding well. Cover and seal<br />

tightly with adhesive film and mix well but carefully by agitation on a plate shaker.<br />

4. Place plate into thermal cycler and run the following program: 94°C for 5 min (94°C for 15 s,<br />

55°C for 15 s, 72°C for 30 s × 35 cycles), and 72°C for 10 min.<br />

5. When cycles are complete, add 4 µL of loading buffer to each sample well and agitate<br />

to mix.<br />

6. Load samples (approx 20 µL) into the wells of an 8% polyacrylamide gel in the vertical<br />

electrophoresis unit and electrophorese at 250 volts for approx 1.25 to 1.5 h along with a<br />

strategically placed molecular weight marker.<br />

7. Remove gel from tank and stain for 2 to 5 min in discarded buffer from upper tank<br />

containing 20 µL of ethidium bromide solution.<br />

8. Interpretation.<br />

The amplification generates three fragments: Prothrombin (F.II), 340 bp; Factor IX<br />

(control), 250 bp; and Factor V, 174 bp. Each set of two tubes, wild type and mutant,<br />

will show bands in the pattern below depending upon the allele detected.<br />

Wild type<br />

Mutant<br />

Homozygous negative –<br />

Heterozygous – –<br />

Homozygous positive –<br />

The Factor IX control fragments must be present in both the wild and mutant tubes<br />

to interpret the P20210A and V Leiden status. Specimens should be reanalyzed if be


ARMS PCR 325<br />

repeated if (1) the Factor IX control bands are not present; (2) other control specimens<br />

are not as expected; (3) the result is homozygous/heterozygous positive; (4) any dubiety<br />

exists with the overall interpretation.<br />

4. Notes<br />

1. Extreme care should be exercised to prevent the possibility of cross-contamination of<br />

samples with amplified DNA.<br />

2. PCR setup should be performed in a physically separated area from product analysis.<br />

3. Samples of known genotype (heterozygous and homozygous negative) should be analyzed<br />

along with unknowns.<br />

References<br />

1. Newton, C. R., Graham, A., Heptinstall, L. E., Powell, S. J., Summers, C., Kalsheker, N.,<br />

et al. (1989) Analysis of any point mutation in DNA. The amplification refractory mutation<br />

system (ARMS). Nucleic Acids Res. 17, 2503–2516.


326 Stirling


PCR-SSCP Analysis of Polymorphism 327<br />

48<br />

PCR-SSCP Analysis of Polymorphism<br />

A Simple and Sensitive Method for Detecting Differences<br />

Between Short Segments of DNA<br />

Mei Han and Mary Ann Robinson<br />

1. Introduction<br />

Polymerase chain reaction single-strand conformation polymorphism (PCR-SSCP)<br />

(1) is a simple method that allows one to rapidly determine whether there are sequence<br />

differences between relatively short stretches of DNA. Coupled with sequence analysis,<br />

SSCP is an extremely useful method for both identifying and characterizing genetic<br />

polymorphisms and mutations. The theory of SSCP is that the primary sequence and<br />

the length of a single stranded DNA fragment determine its conformation when it is<br />

resolved in a nondenaturing polyacrylamide gel. Even single-base differences can cause<br />

different secondary conformations and thus result in different migration rates of the<br />

DNA strands. Radioactive nucleotides are incorporated into the DNA strands by PCR,<br />

making it possible to detect the DNA by autoradiography. SSCP has been widely used<br />

to identify mutations in host genes such as p53 (2–5) and in viruses, such as simian<br />

immuno-deficiency virus (SIV), during the course of infection (6). SSCP has been used<br />

to identify and characterize polymorphisms in a variety of genes (7–9) and was effective<br />

in characterizing alleles of linked genes present in individual sperm (10).<br />

Orita and colleagues (11) developed the SSCP method in 1989 and since then, it has<br />

been applied to screen for sequence differences in either genomic or complementary<br />

DNA (cDNA) samples. Figure 1 schematically shows the steps of the procedure. First,<br />

the region of interest (the gene or cDNA) can be PCR amplified by using primers<br />

corresponding to the desired sequence. Because migration differences are better<br />

resolved using shorter DNA fragments, success is more likely with primers selected<br />

to amplify fragments in 100- to 300-bp range. The PCR mixture contains 32 P-dCTP<br />

to radioactively label the amplification product, which is important for visualizing the<br />

migration differences in later steps of the procedure. The second step is to heat diluted<br />

amplified samples, which will denature the double-stranded DNA into single-stranded<br />

DNA. The samples are mixed with loading buffer containing formamide to hinder<br />

reannealing of the DNA and dye to visually follow the migration of samples through<br />

the gel. The third step is to resolve the single-stranded DNA samples by nondenaturing<br />

polyacrylamide gel electrophoresis. The length of time necessary for running the<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

327


328 Han and Robinson<br />

Fig. 1. Basic steps in SSCP Procedure. 1. The DNA or cDNA fragment of interest is amplified<br />

by PCR incorporating 32 P-dCTP. 2. The PCR amplified-labeled DNA fragment is denatured<br />

by heating. 3. Fragments are separated on an acrylamide gel. 4. Patterns are revealed by<br />

autoradiography.<br />

gel is variable, dependent upon characteristics of the sequence amplified, and can be<br />

determined empirically. The final step is autoradiography to visualize the bands of<br />

radiolabeled DNA in the gel.<br />

Analysis of SSCP autoradiograms is performed visually. Each sequence amplified<br />

in step one will often result in two bands on a corresponding autoradiogram. One band


PCR-SSCP Analysis of Polymorphism 329<br />

Fig. 2. Sample SSCP pattern. The human CD1a gene has two alleles. The DNA sample in<br />

band A derived from an individual homozygous for allele 1, in lane B from a heterozygous<br />

individual, and in lane C from an individual homozygous for allele 2. In this example, one of the<br />

DNA strands assumes two different conformations as it migrates through the gel and results in two<br />

bands. Notice the intensity of the top band compared with the lower bands showing that the top<br />

band corresponds to one strand in one conformation and the lower bands to two conformations<br />

for one strand. The pattern of the heterozygote is a composite of the homozygous patterns. The<br />

single-stranded DNA patterns are shown in this figure. Double-stranded DNA corresponding to<br />

nondenatured PCR product would be found migrating more rapidly on the gel.<br />

corresponds to the upper strand of DNA and the other the lower strand of the amplified<br />

fragment. It is also possible for a segment of single-stranded DNA to assume two<br />

different conformations as it migrates through the gel, and this will result in two bands<br />

for that strand of DNA on a SSCP autoradiogram. Examples of such SSCP patterns are<br />

shown in Fig. 2. There are two alleles of the human CD1a gene (9). DNA samples in<br />

lane A derives from an individual homozygous for allele 1, in lane C, the DNA is from<br />

an individual homozygous for allele 2, and in lane B the DNA is from an individual<br />

heterozygous for alleles 1 and 2. The bands corresponding to alleles 1 and 2 migrate<br />

at different rates, and each has one strand that assumes two different conformations.<br />

The pattern for a heterozygous individual is the composite of the two individual<br />

allele’s patterns.<br />

It was possible to make determinations about the assignment of CD1a alleles because<br />

the analysis was coupled with sequence analysis. To obtain material for sequencing,<br />

the same reaction mixture as used for the SSCP procedure is amplified using an<br />

amplification buffer that does not contain radioactivity for cloning. Many systems are<br />

commercially available for cloning and sequence analysis. In addition, SSCP bands<br />

can be excised from the gel and re-PCR amplified using the same set of primers, which<br />

provides a very convenient way to directly sequence the altered gene segment.<br />

A successful approach to screen for genetic polymorphisms has been to examine<br />

DNA samples from 10 unrelated individuals of diverse ethnic backgrounds. Genetic<br />

polymorphism is likely to be found in such a sampling; however, this does depend upon<br />

allele frequencies and distributions. Although SSCP is quite efficient in identifying


330 Han and Robinson<br />

sequence differences between two segments of DNA, there are cases where differences<br />

are difficult to detect. Changing primers may make a difference. It is possible that the<br />

specific fragment will not show migration differences even if there are substitutions<br />

present. Analysis of overlapping fragments of the region of interest is one solution<br />

to circumvent this difficulty.<br />

The techniques required for the SSCP method are not complicated, and the whole<br />

procedure is time saving (only one-round PCR is needed) compared with other methods<br />

for detecting mutations and DNA polymorphisms. Because of its simplicity and<br />

reliability, PCR-SSCP has been used extensively for identifying alleles and genotypes<br />

in both basic and clinical investigations.<br />

2. Materials<br />

1. Thermal cycling (PCR) machine.<br />

2. AmpliTaq ® DNA polymerase, 5 U/µL (stored at –20°C).<br />

3. 10× PCR buffer stored at –20°C): 100 mM Tris-HCl, pH 8.3, 500 mM KCl, 15 mM<br />

MgCl 2 .<br />

4. dNTP (1.25 mM) stored at –20°C.<br />

5. Primers diluted to a concentration of 20 µM (see Note 1).<br />

6. [α 32 P]-dCTP (New England Nuclear, Boston, MA, BLU013H, 3000 Ci/mmol).<br />

7. DNA templates (100 ng/µL genomic or 110 diluted cDNA).<br />

8. 0.1% SDS and 10 mM EDTA, pH 8.0.<br />

9. Loading buffer: 10 mM NaOH, 95% formamide, 0.05% bromphenol blue, 0.05% xylene<br />

cyanol.<br />

10. Long Ranger 50% stock gel solution (FMC <strong>Bio</strong>Products, Rockland, ME; see Note 2).<br />

11. 10× TBE buffer: 0.89 M Tris-borate; 0.02 M EDTA, pH 8.0.<br />

12. TEMED.<br />

13. 10% Ammonium Persulphate (APS, after dissolving in water, store at –20°C).<br />

14. Glycerol (Ultra Pure, Life Technologies, Gaithersburg, MD).<br />

15. Sigmacote (Sigma Chemical Co., St. Louis, MO).<br />

16. Acrylamide gel electrophoresis supplies and equipment (see Note 3).<br />

17. Whatman 3MM chromatography paper.<br />

18. Plastic wrap (such as Saran Wrap).<br />

19. Vacuum gel dryer.<br />

20. Gel cassette with an enhancing screen.<br />

21. X-ray films.<br />

22. Lab safety area for working with radioactive materials (see Note 4).<br />

3. Methods<br />

3.1. PCR Amplification (see Note 5)<br />

The PCR should contain the following ingredients (in µL): 10× PCR Buffer (2.0);<br />

1.25 mM dNTP (3.2); 20 µM 5′ primer (1.0); 20 µM 3′ primer (1.0); 5 U/µL AmpliTaq ®<br />

(0.1); Experimental template (2.0); [ 32 P]-dCTP (0.1); and H 2 O (10.6) for a total<br />

of 20 µL.<br />

PCR amplification is performed for 30 cycles (denaturation at 95°C for 30 s, annealing<br />

at 55°C (see Note 6) for 30 s, extension at 72°C for 90 s) with a final extension at<br />

72°C for 7 min. PCR amplified DNA may be kept at 4°C or stored at –20°C.


PCR-SSCP Analysis of Polymorphism 331<br />

3.2. SSCP Gel Preparation<br />

1. Prepare a 6% nondenaturing polyacrylamide gel with 10% glycerol as following: Long<br />

Ranger gel solution (7.5 mL), glycerol (7.5 mL), 10× TBE (4.5 mL), H 2 O (55.7 mL);<br />

10% APS (0.4 mL), and TEMED (0.04 mL) for a total of 73.64 mL. Add TEMED and<br />

APS right before pouring the gel.<br />

2. Place a few drops of Sigmacote on the inner side of the shorter plate (see Note 3) and<br />

spread evenly with a paper towel. Allow to air dry. Sigmocote forms a tight microscopically<br />

thin film of silicone on glass, which prevents the SSCP gel from sticking to this particular<br />

plate when you separate the two plates after the gel run is done, allowing the gel to remain<br />

smooth on the other plate. Set up the plates with 0.4-mm spacers on the sides and with<br />

a strip of Whatman chromatography paper at the bottom. Clamp three sides of the plates<br />

leaving the top open.<br />

3. Immediately before pouring the gel, add TEMED and APS to the mixture, collect it into a<br />

60-mL syringe, and push the solution into the plates slowly while holding the plates at an<br />

angle (see Note 7). Place either a 64-well or 36-well comb (see Note 8) into the top of the<br />

gel and clamp the top of the plates. The gel will be polymerized in 2 h.<br />

3.3. Dilution and Denaturation of PCR Products (see Note 9)<br />

1. Dilute 5 µL of amplified DNA in 45 µL of 0.1% SDS and 10 mM EDTA.<br />

2. Mix 5 µL of above diluted PCR products with 5 µL of loading buffer. The sample is ready<br />

to load and may be stored at –20°C for 2 wk.<br />

3.4. Loading Samples and Running Gels<br />

1. Before running the gel, denature the diluted PCR samples by heating at 95°C for 2 min<br />

and cool rapidly in an ice bath.<br />

2. Load 2 µL (for 64-well combs) or 5 µL (for 32-well combs) of the diluted PCR products<br />

onto gel. The leftover samples can be reused if they have been stored at –20°C for less<br />

than 2 wk.<br />

3. Electrophoresis is performed by using 0.6× TBE buffer at a constant 30 watts (W) for<br />

4 h at room temperature (see Note 10).<br />

3.5. Drying Gels and Developing Films<br />

1. After migration, the gel is transferred onto a double layer of Whatman Chromatography<br />

paper and covered with plastic wrap avoiding air bubbles.<br />

2. The gel is vacuum dried in a gel dryer at 80°C for 1 h, and then the gel is placed facing<br />

up in a film cassette.<br />

3. Expose gel to an X-ray film (the X-ray film should be put in between the enhancing screen<br />

and the gel) at –80°C for 16 to 20 h (see Note 11).<br />

4. Warm up film or air dry film before developing. Cut a corner of the film before developing,<br />

which will help to figure out the sample orders to load. Develop film using an automatic<br />

film developer or using the dip method.<br />

3.6. SSCP Gel Analysis<br />

SSCP patterns are analyzed upon visual examination. Analysis is facilitated with<br />

knowledge of the gene sequence, which is often available because sequence <strong>info</strong>rmation<br />

was necessary to derive primers. Inclusion of control samples of known sequence allow<br />

allele typing (see Note 12 and Subheading 1.).


332 Han and Robinson<br />

4. Notes<br />

1. PCR-SSCP is a very sensitive method for detecting point mutations in a DNA fragment<br />

shorter than 300 bp. A single base change may effect the overall conformation more easily<br />

on small fragments than large fragments. Therefore, primers used in your experiment are<br />

better designed for amplifying a 100- to 300-bp DNA fragment. If long genomic or cDNA<br />

segments are to be screened, it is important to find comparable electrophoresis conditions<br />

in which mutations in long DNA fragments are efficiently detected (see Note 10). If the<br />

primers used do not allow detection of known substitutions, moving the primer sequence<br />

further 5′ or 3′ may make it possible to distinguish the fragments. The sequences of the<br />

oligo primers used to amplify the human CD1a gene, as shown in Fig. 2, are as follows:<br />

CD1a.ex2F1 AGACGGGCTCAAGGAGCCTC and CD1a.ex2R1 TCCAGTTCCTTC<br />

CACTCCTC.<br />

2. Long Ranger stock gel solution contains acrylamide, a neurotoxin, which may cause cancer<br />

and /or heritable genetic damage. It is advisable to wear gloves while handling.<br />

3. Acrylamide gel electrophoresis supplies and equipment format used in PCR-SSCP method<br />

may vary between laboratories. In our laboratory, we use a set of glass plates (44.5 × 37.5 cm<br />

and 42 × 37.5 cm) and corresponding electrophoresis supplies. Mini-gel format is also<br />

frequently used by some other investigators and which is often followed by silver-staining<br />

DNA bands rather than labeling with 32 P.<br />

4. All procedures involved in dealing with [ 32 P]-dCTP should be performed in a lab safety<br />

area and handled by behind a protection screen. Radioactive waste should only be thrown<br />

into specialized trash cans.<br />

5. The final concentration of 1 µM primer in our PCR amplification provides clear migration<br />

bands, and we have never failed to detect known mutations using this primer concentration.<br />

However, it has been reported that lowering the upstream primer concentration improves<br />

DNA migration. Because the SSCP technique is not 100% effective in detecting any given<br />

substitution, it might be worthwhile to try lowering upstream primer concentrations if<br />

known base changes are not detected in your SSCP gel. The template can either be genomic<br />

or cDNA depending upon your interest. But do not forget to include DNA samples with<br />

known sequence in your PCR amplification, which will serve as a control SSCP pattern<br />

for reading allelic distributions of your test samples.<br />

6. Annealing temperature may vary depending upon the primers used. Usually the annealing<br />

temperature is determined empirically with estimates made by the following formula:<br />

[(G + C) × 4] °C + [(A + T) × 2] °C = annealing temperature °C.<br />

7. Some bubbles may appear between the plates if the plates are not very clean. You can get<br />

the bubbles out by further angling the plates with one hand and tapping on the plates with<br />

the other hand until the bubbles are out.<br />

8. It is important to ensure that the comb is very clean when casting the gel. Leftover<br />

acrylamide present between the teeth of the comb may cause shorter wells or a gel surface<br />

that will not hold your samples.<br />

9. Denaturation of amplified DNA samples can be performed by heating at 95°C or by both<br />

chemical agents (NaOH) and heating because the denaturation of the fragment to single<br />

strands is often incomplete.<br />

10. The optimum balance for band separation versus gel run time depends upon the size<br />

of amplified DNA fragment. For a DNA fragment around 150 to 300 bp, 4 h would be<br />

ideal. The migration pattern of a certain single-strand conformer may be improved by<br />

modification of a number of electrophoresis conditions, including lowering the buffer pH,<br />

changing the temperature of the gel, or increasing the acrylamide concentration (up to<br />

15%). Difficulties detecting certain mutations when using the routine protocol may be<br />

circumvented by adjusting one or more of these three gel conditions.


PCR-SSCP Analysis of Polymorphism 333<br />

11. Usually the SSCP patterns can be read clearly after 16- to 20-h exposure. If the migration<br />

bands are still very weak after exposed for 20 h, exposure time should be prolonged to<br />

more than 48 h, sometimes even for as long as 2 wk.<br />

12. SSCP gels are analyzed by observing the pattern of bands resulting on the autoradiograms.<br />

The number of bands present is dependent upon the sequence amplified. When one<br />

sequence is amplified, the most simple pattern that can be observed consists of two<br />

bands, one corresponding to the upper strand of DNA, and the other to the lower strand<br />

of the amplified fragment. If a segment of single-stranded DNA assumes more than one<br />

conformation as it migrates through the gel, the result will be multiple bands (equal<br />

to the number of conformers). Knowledge of the sequence being amplified provides a<br />

control and a point of reference. Amplification of more than one sequence results in a<br />

composite pattern of bands corresponding to those of the constituent sequences. A band<br />

that corresponds to double-stranded DNA may also be visible lower in the gel as doublestranded<br />

DNA migrates more rapidly. Inclusion of a sample of nondenatured DNA (not<br />

heated) serves as a marker for double-stranded DNA.<br />

References<br />

1. Orita, M., Suzuki, Y., Sekiya, T., and Hayashi, K. (1989) Rapid and sensitive detection of<br />

point mutations and DNA polymorphisms using the polymerase chain reaction. Genomics<br />

5, 874–879.<br />

2. Bosari, S., Marchetti, A., Buttitta, F., Graziani, D., Borsani, G., Loda, M., et al. (1995)<br />

Detection of p53 mutations by single-strand conformation polymorphisms (SSCP) gel<br />

electrophoresis. A comparative study of radioactive and nonradioactive silver-stained SSCP<br />

analysis. Diagn. Mol. Pathol. 4, 249–255.<br />

3. Kurvinen, K., Hietanen, S., Syrjanen, K., and Syrjanen, S. (1995) Rapid and effective<br />

detection of mutations in the p53 gene using nonradioactive single-strand conformation<br />

polymorphism (SSCP) technique applied on PhastSystem. J. Virol. Methods 51, 43–53.<br />

4. Neubauer, A., Brendel, C., Vogel, D., Schmidt, C. A., Heide, I., and Huhn, D. (1993)<br />

Detection of p53 mutations using nonradioactive SSCP analysis: p53 is not frequently<br />

mutated in myelodysplastic syndromes (MDS). Ann. Hematol. 67, 223–226.<br />

5. Ozcelik, H. and Andrulis, I. L. (1995) Multiplex PCR-SSCP for simultaneous screening for<br />

mutations in several exons of p53. <strong>Bio</strong>techniques 18, 742–744.<br />

6. Campbell, B. J. and Hirsch, V. M. (1994) Extensive envelope heterogeneity of simian<br />

immunodeficiency virus in tissues from infected macaques. J. Virol. 68, 3129–3137.<br />

7. Barron, K. S., Reveille, J. D., Carrington, M., Mann, D. L., and Robinson, M. A. (1995)<br />

Susceptibility to Reiter’s syndrome is associated with alleles of TAP genes. Arthritis<br />

Rheum. 38, 684–689.<br />

8. Barron, K. S. and Robinson, M. A. (1994) The human T-cell receptor variable gene segment<br />

TCRBV6S1 has two null alleles. Hum. Immunol. 40, 17–19.<br />

9. Han, M., Hannick, L. I., DiBrino, M., and Robinson, M. A. (1999) Polymorphism of human<br />

CD1 genes. Tissue Antigens 54, 122–127.<br />

10. Day, C. E., Schmitt, K., and Robinson, M. A. (1994) Frequent recombination in the human<br />

T-cell receptor beta gene complex. Immunogenetics 39, 335–342.<br />

11. Orita, M., Iwahana, H., Kanashi, H., and, Sekiya, T. (1989) Detection of polymorphisms of<br />

human DNA by gel electrophoresis as single-strand conformation polymorphisms. Proc.<br />

Natl. Acad. Sci. USA 86, 2766–2770.


334 Han and Robinson


Sequencing 337<br />

49<br />

Sequencing<br />

A Technical Overview<br />

David Stirling<br />

1. Introduction<br />

The huge advances that have been made in human (and other species) genome<br />

projects have been lead by, and have themselves fueled, tremendous innovations in<br />

the field of sequencing. These innovations have lead to a specialization, with very<br />

expensive equipment, often in core sequencing centers or services. Until the cost of<br />

automated sequencing equipment falls considerably, the economics of scale mean that<br />

unless an investigator has a huge amount of sequencing to perform, they are most likely<br />

to be best served by using these services rather than by investing in the technology<br />

themselves. There is always a temptation in such situations to treat the sequencing as a<br />

black box, without worrying too much how the data are generated. This is a mistake. A<br />

clear understanding of the principals involved will not only aid the design of sequencing<br />

strategies but will also help in the interpretation of less than perfect data.<br />

DNA/RNA sequence analysis has become fundamental to the understanding of<br />

biological processes. The foundations of this science were established in 1977 when<br />

Maxam and Gilbert (1) described a method for sequencing by base-specific chemical<br />

cleavage. Subsequently, Sanger and co-workers (2) developed a method for enzymatic<br />

sequencing using chain terminators. Both techniques produce populations of labeled<br />

DNA fragments that can be electrophoretically resolved to reveal the base sequence.<br />

Many different strategies have been developed to improve on these initial approaches<br />

and to make genome-sequencing projects feasible; all owe a great debt to those original<br />

protocols.<br />

2. Chemical DNA Sequencing<br />

In the original Maxam-Gilbert method of DNA sequencing, target DNA is radioactively<br />

labeled at one end (3′or 5′ end), which acts as the reference point for determining<br />

the positions of the remaining bases. This labeled DNA is processed with four basespecific<br />

reactions. First, a base-specific chemical modification, then a chain cleavage<br />

of modified bases.<br />

These methods have fallen from popularity, both because of the toxicity of the<br />

reagents (dimethyl sulphate, hydrazine, potassium permanganate) and the development<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

337


338 Stirling<br />

of simple and improved methods for enzymatic DNA. Although the chemical method<br />

is not widely used as the enzymatic method, it has some advantages and can be very<br />

useful in certain situations. Because it does not rely on the hybridization of a primer,<br />

very short sequences, such as oligonucleotides, can be analyzed. It is also useful<br />

for analyses of DNA modifications, such as methylation, and to study DNA–protein<br />

interactions (footprinting).<br />

3. Enzymatic DNA Sequencing<br />

Dideoxy sequencing reactions (Sanger method) are essentially primer extension<br />

reactions where ddNTPs are included in the mix. In the conventional dideoxy sequencing<br />

reaction, an oligo primer is annealed to a single-stranded DNA template and<br />

extended by DNA polymerase to synthesize a complementary copy of single-stranded<br />

DNA in the presence of four dNTPs, one of which is 35 S-labeled. Chain growth involves<br />

the formation of a phosphodiester bridge between the 3′-OH at the growing end of<br />

the primer and the 5′-phosphate group of the incorporated dNTP. Thus, overall chain<br />

growth is in the 5′→3′ direction. The reaction also contains one of four ddNTPs that<br />

terminate elongation when incorporated into the growing DNA chain. When a ddNTP<br />

is incorporated at the 3′ end of the growing primer chain, the elongation is terminated<br />

selectively at A, C, G, or T owing to the missing 3′ OH group of the primer chain. The<br />

enzymatic method is based on the ability of DNA polymerase to use both 2′ dNTPs and<br />

2′,3′ ddNTPs as substrates. After completion of the sequencing reactions, the products<br />

are subjected to electrophoresis on a high-resolution denaturing polyacrylamide gel<br />

and then autoradiographed to visualize the DNA sequence.<br />

For this form of sequencing, the purity and concentration of the template has to be<br />

very carefully controlled, and the template has to be rendered single stranded before<br />

the reaction will proceed. To reliably obtain good quality sequence from PCR products,<br />

these would first have to be cloned into a suitable vector. Any clone obtained is the<br />

product of a single molecule of PCR product, and so the likelihood of that sequence<br />

containing misincorporated bases is relatively high. Multiple clones have therefore to<br />

be sequenced to produce reliable data.<br />

4. Cycle Sequencing<br />

The introduction of thermostable DNA polymerase had the same revolutionary<br />

impact on sequencing protocols as in other areas (3). The ability to function at higher<br />

temperature resulted in template remaining single stranded for longer periods, overcame<br />

many problems of secondary structure, and allowed more stringent primer annealing<br />

conditions resulting on less background noise in sequencing data. In addition, repeated<br />

cycles of primer annealing and extension amplifies (linearly rather than exponentially as<br />

for PCR) even small amounts of sample DNA to generate more template. The ability to<br />

sequence very low template concentrations means that even relatively impure material,<br />

such as PCR products, can be used, simply by diluting out the impurities.<br />

5. Automation<br />

One of the major advances in sequencing technology has been the development of<br />

automated DNA sequencers, which automate the gel electrophoresis step, detection of<br />

band pattern, and analysis of bands. These machines are based on the enzymatic, cycle


Sequencing 339<br />

sequencing approach and use fluorescent rather than radioactive labels. This has the<br />

advantages of greater safety, generation of machine-readable data, and greater reagent<br />

stability. The downside of this is the relative expense of the equipment required.<br />

Fluorescent dyes can be incorporated as labeled primers, dNTPs. or ddNTPs. The use<br />

of four different dyes, one for each of the ddNTPs (dye terminator sequencing), has<br />

allowed one of the biggest improvements in throughput of these systems. Now, rather<br />

than the base-specific reactions being kept separate and run down individual lanes<br />

of a gel, they can be performed in the same reaction tube and analyzed in one lane,<br />

with fragments being separated by size and distinguished by the wavelength of the<br />

fluorescent emission. Automated sequencers based on capillary electrophoresis have<br />

been developed, which dispense with the gel-making step. Many specialized centers<br />

now use robotic systems to extract DNA template, perform PCR and sequencing reactions,<br />

and load 96-capillary sequencers. Sequence data generated can then be directly<br />

imported into databases and processed with little or no hands-on intervention.<br />

These developments still continue. Research is under way to develop the technology<br />

of mass spectrometry for DNA sequencing, and sequencing by hybridization is the<br />

subject of a great deal of development work.<br />

6. Direct Sequencing of PCR Products<br />

Unlike methods where the PCR product is cloned and a single clone is sequenced,<br />

direct sequencing of PCR products is usually unaffected by the relatively high error<br />

rate of Taq DNA polymerase because (unless there are only a few starting copies of<br />

template, and a misincorporation occurs in an early round of PCR) the vast majority of<br />

the amplified product will consist of the correct sequence.<br />

Direct sequencing of PCR products has significant advantages over the cloning<br />

strategy. It is a simple procedure that can be easily standardized and only a single<br />

sequence needs to be determined for each sample. Indeed, the procedures are so well<br />

standardized that there has been an exponential growth in the number of laboratories<br />

offering core-sequencing services for PCR products. Although the majority of such<br />

core laboratories provide an excellent service, there are a number of factors that will<br />

affect their ability to produce good quality data.<br />

7. Requirements for Good Quality Sequence from PCR Products<br />

• Optimize the PCR reaction to yield only a single product. If the same oligos are used<br />

to sequence as primed the PCR, they will also sequence from any nonspecific products,<br />

interfering with data quality. Primer-dimers may not interfere with the identification of<br />

a PCR product on agarose gel electrophoresis, but they serve as efficient template for<br />

sequencing, resulting in characteristic noise in sequence data close to the primer.<br />

• PCR should be performed with as little primer as possible. If labeled dideoxyterminator<br />

sequencing is used (probably the most common approach), residual PCR primer can serve<br />

as sequencing primer. Thus, even when only one PCR product is present, it is sequenced<br />

from each strand simultaneously, again interfering with data quality. There are a number of<br />

techniques available for the removal of unincorporated primers, from enzymatic digestion<br />

to column purification, but these are generally unnecessary if the primer concentration<br />

is limited in the PCR.<br />

• If PCR products are isolated from a gel prior to sequencing, great care should be taken<br />

to minimize the amount of salt carried over from the isolation procedure. High-salt


340 Stirling<br />

concentrations interfere with the processivity of the polymerase reactions, resulting in<br />

very short read lengths.<br />

• Primers should only be obtained from reliable sources and should be aliquoted to avoid<br />

repeated freeze thaw cycles. Any shorter primer species (e.g., n-1), in the PCR primer may<br />

not affect PCR efficiency. However, if the same primer is used for sequencing, a proportion<br />

of the fluorescent products will be primed from a shorter oligo. If the missing base is at the<br />

5′ end of the oligo, the product will be shorter, and hence the sequencing results will be<br />

confounded. This often results in a characteristic ‘shadow’ sequence.<br />

References<br />

1. Maxam, A. M. and Gilbert, W. (1977) A new method for sequencing DNA. Proc. Natl.<br />

Acad. Sci. USA 74, 560–564.<br />

2. Sanger, F., Nicklen, S., and Coulson, A. R. (1977) DNA sequencing with chain-terminating<br />

inhibitors. Proc. Natl. Acad. Sci. USA 74, 5463–5467.<br />

3. Innis, M. A., Myambo, K. B., Gelfand, D. H., and Brow. M. A. (1988) DNA sequencing<br />

with Thermus aquaticus DNA polymerase and direct sequencing of polymerase chain<br />

reaction-amplified DNA. Proc. Natl. Acad. Sci. USA 85, 9436–9440.


Direct Automated Cycle Sequencing 341<br />

50<br />

Preparation and Direct Automated Cycle Sequencing<br />

of PCR Products<br />

Susan E. Daniels<br />

1. Introduction<br />

The polymerase chain reaction (PCR) is well known for being a rapid and versatile<br />

method for the amplification of defined target DNA sequences. This technique can be<br />

applied to a variety of research areas, such as the identification and typing of single<br />

nucleotide substitutions of DNA sequence polymorphisms, and genetic mapping (1–4).<br />

Since the introduction of PCR (5), a variety of methods for sequencing PCRgenerated<br />

fragments have been described. These are usually based on the Sanger<br />

chain-terminating dideoxynucleotide sequencing (6) rather than the Maxam and Gilbert<br />

chemical cleavage method (7). Manual dideoxy sequencing methods are labor intensive,<br />

time-consuming, involve radioisotopes, and have limitations in sequence ordering.<br />

However, a technique combining the PCR and dideoxy terminator chemistry simplifies<br />

the process of sequencing and is known as cycle sequencing (8). Automated or<br />

fluorescent DNA sequencing is a variation of the traditional Sanger sequencing<br />

using the cycle sequencing methodology, where fluorescent labels are covalently<br />

attached to the reaction products and data are collected during the polyacrylamide<br />

gel electrophoresis.<br />

The introduction of fluorescently labeled dideoxynucleotides as chain terminators<br />

presented the opportunity for the development of reliable cycle sequencing for PCR<br />

products. The sequencing reaction with the dye terminators is performed in a thermal<br />

cycler, and each of the four dideoxynucleotide triphosphates (ddNTPs) is labeled<br />

with a different fluorescent dye. This allows the four chain extension reactions to be<br />

conducted within a single tube, sparing considerable labor (9,10). The use of labeled<br />

chain terminators allows flexibility of sequencing strategy because the same primers<br />

can be used in the sequencing reaction. This eliminates the time and expense associated<br />

with a separate set of modified DNA sequencing primers and is well suited to high<br />

throughput sequencing.<br />

Using this method, it is possible to amplify a target DNA sequence, purify the<br />

resulting fragment, and obtain sequencing data within 24 h. Also, it has been used<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

341


342 Daniels<br />

to sequence a 500-bp PCR fragment on an automated DNA sequencer with 99.3%<br />

accuracy (10–12).<br />

This chapter focuses on the techniques involved with the direct sequencing reactions<br />

rather than on the use of the machine because each automated DNA sequencer will be<br />

provided with an extensive manual for its operation.<br />

2. Materials<br />

All solutions should be made to the standard required for molecular biology. Use<br />

molecular-biology-grade reagents and sterile distilled water. The reagents for the cycle<br />

sequencing are available commercially.<br />

2.1. Purification of PCR Products Before Cycle Sequencing<br />

1. Ammonium acetate (4 M).<br />

2. Isopropanol.<br />

3. 70% (v/v) ethanol.<br />

4. Tris-HCl (10 mM, pH 7.5), 1 mM EDTA.<br />

2.2. Cycle Sequencing<br />

Prism ready reaction DyeDeoxy terminator premix (1000 µL, Applied <strong>Bio</strong>systems<br />

[ABI]) consists of 1.58 mM A-dyedeoxy, 94.74 µM T-dyedeoxy, 0.42 µM G-dyedeoxy,<br />

47.37 µM C-dyedeoxy, 78.95 µM dITP, 15.79 µM dATP, 15.79 µM dCTP, 15.79 µM<br />

dTTP, 168.42 mM Tris-HCl (pH 9.0), 4.21 mM (NH 4 ) 2 SO 4 , 42.1 mM MgCl 2 , 0.42 U/µL<br />

AmpliTaq DNA polymerase.<br />

2.3. Purification of PCR Products After Cycle Sequencing<br />

1. Chloroform.<br />

2. PhenolH2Ochloroform (161814) at room temperature.<br />

3. Sodium acetate (2 M, pH 4.5).<br />

4. 100 and 70% (v/v) ethanol at room temperature.<br />

2.4. 6% Polyacrylamide Sequencing Gels<br />

1. 10× TBE: 890 mM Tris-borate, 890 mM boric acid, and 20 mM EDTA, pH 8.3.<br />

2. Urea (40 g).<br />

3. 12 mL of 40% (w/v) acrylamide stock solution (191 acrylamide/bis-acrylamide).<br />

4. dH 2 O (20 mL).<br />

5. Mixed-bed ion-exchange resin (1 g).<br />

6. TEMED.<br />

7. 10% (w/v) ammonium persulfate, freshly made.<br />

2.5. Loading Buffer<br />

1. 50 mM EDTA, pH 8.0.<br />

2. Deionized formamide.<br />

3. Methods<br />

3.1. Isopropanol Purification of PCR Products<br />

It is essential to remove excess PCR primers before using DyeDeoxy terminators<br />

for cycle sequencing (see Note 1).


Direct Automated Cycle Sequencing 343<br />

1. Aliquot an appropriate amount of the PCR into a 0.6-mL microfuge tube, and dilute to<br />

a total of 20 µL with distilled water.<br />

2. Add 20 µL of 4 M ammonium acetate into the microfuge tube, mixing well.<br />

3. Add 40 µL of isopropanol into the tube, mix well, leave at room temperature for 10 min,<br />

and centrifuge the microfuge tube for 10 min at 12,000g.<br />

4. Carefully remove the supernatant and wash the pellet with 70% (v/v) ethanol. Then, briefly<br />

dry the pellet under vacuum.<br />

5. Resuspend the pellet in 20 µL of TE buffer.<br />

3.2. Cycle Sequencing of PCR Products<br />

The amount of PCR product should be estimated on an agarose gel before sequencing.<br />

Approximately 1 µg of double-stranded DNA template or 0.5 µg of single-stranded<br />

DNA template is required for each sequencing reaction.<br />

1. Mix the following reagents in a 0.6-mL microfuge tube: 5 mL of DNA template, 1 µL<br />

of primer (from a 3.2-pmol stock solution), and 4.5 µL of sterile dH 2 O (see Notes 4<br />

and 5).<br />

2. Add 9.5 µL of ABI Prism ready reaction DyeDeoxy terminator premix.<br />

3. Spin briefly to collect the reaction mix in the bottom of the tube and overlay with approx<br />

50 µL of mineral oil.<br />

4. Place the tubes in a thermal cycler (Perkin–Elmer Cetus [Warrington, UK] model 480 or<br />

9600) that has been preheated to 96°C.<br />

5. Immediately begin the cycle sequencing program, which is as follows: rapid thermal ramp<br />

to 96°C; 96°C for 30 s; Rapid thermal ramp to 50°C; 50°C for 15 s; Rapid thermal ramp<br />

to 60°C; and 60°C for 4 min for a total of 25 cycles.<br />

6. Try to keep the samples in the dark at 4°C until further processing because they are<br />

sensitive to light.<br />

7. Remove the excess DyeDeoxy terminators from the completed sequencing reactions.<br />

3.3. PhenolChloroform Extraction of Cycle Sequencing Products<br />

This step is essential to remove excess primers and unincorporated nucleotides.<br />

1. To each sample, add 80 µL of sterile dH 2 O.<br />

2. Either add 100 µL of chloroform to dissolve the oil or remove the oil with a pipet.<br />

3. Add 100 µL of phenolH 2 Ochloroform (681814) to the sample and mix well by<br />

vortexing.<br />

4. Centrifuge the sample at 12,000g for 1 min. Remove and discard the lower organic phase.<br />

5. Re-extract the aqueous layer and transfer the aqueous upper layer to a clean tube.<br />

6. Add 15 µL of 2 M sodium acetate and 300 µL of 100% ethanol to precipitate the extension<br />

products (see Note 7).<br />

7. Centrifuge at 12,000g for 15 min at room temperature.<br />

8. Carefully remove the supernatant and wash the pellet with 70% ethanol. Then, briefly dry<br />

the pellet under a vacuum (see Note 2).<br />

3.4. Preparation of Samples for Loading<br />

1. Add 4 µL of deionized formamide: 50 mM EDTA, pH 8.0 (51), to each sample tube, and<br />

mix well to dissolve the dry pellet.<br />

2. Centrifuge briefly to collect the liquid at the bottom of the tube.<br />

3. Before loading, heat the samples at 90°C for 2 to 3 min to denature. Then, transfer<br />

immediately onto ice.


344 Daniels<br />

4. Load all of the samples onto the automated DNA sequencer fitted with a 6% polyacrylamide<br />

gel using the manufacturer’s software.<br />

3.5. Preparation of 6% Polyacrylamide Sequencing Gel (see Note 3)<br />

1. Place 40 g of urea, 12 mL of 40% acrylamide stock, 20 mL of dH 2 O, and 1 g of mixedbed<br />

ion-exchange resin into a beaker and stir gently while warming. Continue to stir the<br />

solution until all the urea crystals have dissolved (see Notes 8 and 9).<br />

2. Filter the acrylamide through a 0.2-µM filter, degas for 5 min, and transfer to a 100-mL<br />

cylinder.<br />

3. Add 8 mL of filtered 10× TBE buffer and adjust the volume to 80 mL with dH 2 O.<br />

4. To polymerize the gel, add 400 µL of 10% APS (freshly made) and 45 µL of TEMED.<br />

Gently swirl to avoid adding air bubbles.<br />

5. According to the instructions provided for the automated DNA sequencer, run the<br />

sequencing gel and analyze the readouts as indicated in Fig. 1 (see Note 10).<br />

4. Notes<br />

1. Both purification steps can also be performed by spin columns, such as Centri-Sep or<br />

Quick Spin. Although they provide a quicker purification, they are an expensive alternative<br />

for those on a tight budget.<br />

2. The PCR primers can be used as the DNA sequencing primers, and should be at least an<br />

18 mer in length. Increasing the length will increase specificity and prevents priming at a<br />

secondary site. It will also decrease the chances for nonexact hybridization.<br />

3. The GC content of the primer should be between 50 and 60%.<br />

4. After ethanol or isopropanol precipitation, it is very important that the supernatant be<br />

carefully aspirated because the pellets are unstable and might be lost.<br />

5. The dried sequencing pellet can be stored in the dark at 4°C for several days if required.<br />

However, once the loading buffer has been added, the samples should be loaded within<br />

a few hours.<br />

6. The sequencing gel must polymerize for at least 1 h before use. A good time to prepare<br />

the gel is during the PCR of the cycle sequencing reactions.<br />

7. Although there are various ready-made acrylamide solutions on the market, it is recommended<br />

that you make your own solutions because better resolution will be achieved.<br />

However, 40% acrylamide can be purchased ready-made and gives good results.<br />

8. It has been shown that the use of formamide in the polyacrylamide gels can resolve<br />

compressions (13).<br />

9. During the phenol:chloroform extraction, if after the first spin two separate layers are not<br />

seen, then revortex the samples for 1 min and recentrifuge, after which an aqueous and<br />

organic phase should be obtainable.<br />

10. If using the ABI 373A “Stretch” automated DNA sequencers, then it is better to use a<br />

4.75% polyacrylamide gel, and from these machines, it is possible to obtain up to 1 kb<br />

of sequence with one run.<br />

11. If the thermal cycler in your laboratory is not a Perkin–Elmer Cetus, then it will be<br />

necessary to optimize your thermal cycler for the sequencing reactions.<br />

12. The addition of blue dextran to the loading buffer will make gel loading easier.


Direct Automated Cycle Sequencing 345<br />

Fig. 1. Analyzed sequence data for a 350-bp PCR fragment amplified from genomic human DNA. The data were obtained<br />

using one of the primers used for PCR amplification.<br />

345


346 Daniels<br />

References<br />

1. Orita, M., Suzuki, Y., Sekeiya, T., and Hayashi, K. (1989) Rapid and sensitive detection of<br />

point mutations and DNA polymorphisms using the polymerase chain reaction. Genomics<br />

5, 874–879.<br />

2. Kwok, P. Y., Carlson, C., Yager, T. D., Ankener, W., and Nickerson, D. A. (1994) Comparative<br />

analysis of human DNA variations by fluorescence based sequencing of PCR products.<br />

Genomics 23, 138–144.<br />

3. Martin-Gallardo, A., McCombie, W. R., Gocayne, J. D., Fitzgerald, M. G., Wallace, S.,<br />

Lee, B. M. B., et al. (1992) Automated DNA sequencing and analysis of 106 kilobases<br />

from human chromosome 19q 13.3. Nat. Genet. 1, 34–39.<br />

4. NIH/CEPH Collaborative Mapping Group. (1992) A comprehensive genetic linkage map<br />

of the human genome. Science 258, 67–86.<br />

5. Mullis, K. B. and Faloona, F. A. (1987) Specific synthesis of DNA in vitro via a polymerase<br />

catalysed chain reaction. Methods Enzymol. 155, 335–350.<br />

6. Sanger, F., Nicklen, S., and Coulson, A. R. (1977) DNA sequencing with chain terminating<br />

inhibitors. Proc. Natl. Acad. Sci. USA 74, 5463–5467.<br />

7. Maxam, A. M. and Gilbert, W. (1977) A new method for sequencing DNA. Proc. Natl.<br />

Acad. Sci. USA 74, 560–564.<br />

8. Carothers, A. M., Urlaub, G., Mucha, J., Grunberger, D., and Chasin, L. A. (1989) Point<br />

mutation analysis in a mammalian gene: Rapid preparation of total RNA, PCR amplification<br />

of cDNA, and Taq sequencing by a novel method. <strong>Bio</strong>Techniques 7, 494– 499.<br />

9. McBride, L. J., Koepf, S. M., Gibbs, R. A., Nyugen, P., Salser, W., Mayrand, P. E., et al.<br />

(1989) Automated DNA sequencing methods involving polymerase chain reaction. Clin.<br />

Chem. 35, 2196–2201.<br />

10. Tracy, T. E. and Mulcahy, L. S. (1991) A simple method for direct automated sequencing<br />

of PCR fragments. <strong>Bio</strong>Techniques 11, 68–75.<br />

11. Rosenthal, A. and Charnock Jones, D. S. (1992) New protocols for DNA sequencing with<br />

dye terminators. DNA Sequence 3, 61–64.<br />

12. Kelley, J. M. (1994) Automated dye terminator DNA sequencing, in Automated DNA<br />

Sequencing and Analysis (Adams, M. A., Fields, C., and Venter, J. C., eds.), Academic,<br />

London, pp. 175–181.<br />

13. Hawkins, T. L. and Sulston, J. E. (1991) The resolution of compressions in automated<br />

fluorescent sequencing. Nucleic Acids Res. 19, 2784.


Digoxigenin 347<br />

51<br />

Nonradioactive PCR Sequencing Using Digoxigenin<br />

Siegfried Kösel, Christoph B. Lücking, Rupert Egensperger,<br />

and Manuel B. Graeber<br />

1. Introduction<br />

Techniques for direct sequencing of polymerase chain reaction (PCR) products are<br />

of central importance to contemporary research in molecular biology and genetics. The<br />

rapidly growing number of cloned human disease genes increasingly allows sequencing<br />

of PCR amplicons for diagnostic purposes. Nonradioactive sequencing protocols are of<br />

particular use because health, environmental, and administrative risks are minimized<br />

compared with conventional isotopic methods. The PCR-based nonradioactive cycle<br />

sequencing protocol described in this chapter has been successfully used to sequence<br />

mitochondrial and nuclear genes in Parkinson’s and Alzheimer’s disease brains using<br />

DNA extracted from formalin-fixed and paraffin-embedded neuropathological material<br />

(1–3). This method, which allows sequence <strong>info</strong>rmation of PCR products to be<br />

obtained within a single day, can be performed in a research or clinical laboratory using<br />

relatively inexpensive equipment.After initial PCR amplification, amplicons are purified<br />

using spin columns for affinity chromatography or ultrafiltration. Subsequently, cycle<br />

sequencing (4) is performed using 5′-digoxigenin end-labeled oligonucleotide primers.<br />

Because the nucleotide sequences of the PCR and sequencing primers can be identical,<br />

both reactions may be performed using the same thermalcycling protocol. This obviates<br />

the need for time-consuming optimization procedures. For visualization of sequencing<br />

results, sequencing reactions are separated on a standard sequencing gel, the gel is<br />

contact-blotted to a nylon membrane, and sequencing bands are visualized using<br />

alkaline phosphatase-conjugated antibodies (Fig. 1).<br />

Potential pitfalls of our method are primarily related to the extreme sensitivity of PCR.<br />

The need for positive and negative sample controls cannot be overemphasized. We use<br />

different rooms and different pipets for setting up PCRs, pipetting sequencing templates,<br />

and thermal cycling (5). In addition, aerosol-resistant pipet tips are always used.<br />

2. Materials<br />

2.1. Purification of Sequencing Templates<br />

1. 10× TNE: 100 mM Tris-HCl, pH 7.4, 1.0 M NaCl, 10 mM EDTA (see Note 1).<br />

2. Wizard PCR Preps DNA Purification System containing affinity chromatography spin<br />

columns, purification buffer, and resin (A7170, Promega, Madison, WI). Not contained<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

347


348 Kösel et al.<br />

Fig. 1. Schematic drawing summarizing the essential steps of nonradioactive PCR sequencing<br />

using digoxigenin.<br />

in the kit are 1× TE: 10 mM Tris-HCl, pH 8.0, 1 mM EDTA (6); 80% isopropanol; and<br />

2-mL disposable syringes (one per reaction) or, alternatively, Microcon-30 Concentrators<br />

(#42410, Amicon, Beverly, MA).<br />

3. Eppendorf tubes (15 mL).<br />

4. Eppendorf centrifuge (#5415C Eppendorf, Hamburg, Germany).<br />

5. Fluorometer (TKO 100, Hoefer, San Francisco, CA) for quantification of template<br />

concentrations using DNA dye Hoechst No. 33258 and calf thymus standard DNA<br />

(supplied with the fluorometer).<br />

2.2. Sequencing Reactions<br />

1. Sequencing primer, 5′digoxigenin end-labeled (1 pmol/µL), desalted, and HPLC-purified<br />

(see Note 2).<br />

2. Mineral oil (M-5904, Sigma, St. Louis, MO).<br />

3. Dig Taq DNA Sequencing Kit (#1449443, Boehringer Mannheim, Mannheim, Germany),<br />

containing Taq DNA polymerase (3 U/µL); sequencing buffer (250 mM Tris-HCl,


Digoxigenin 349<br />

pH 8.0, 50 mM MgCl 2 ); loading buffer containing formamide; and Termination mixtures<br />

containing dNTPs and the appropriate ddNTP:<br />

ddATP (dATP, dCTP, dGTP, and dTTP, 25 µM each; 850 µM ddATP; 950 µM MgCl 2 ;<br />

pH 7.5).<br />

ddCTP (dATP, dCTP, dGTP, dTTP, 25 µM each; 400 µM ddCTP; 500 µM MgCl 2 ;<br />

pH 7.5).<br />

ddGTP (dATP, dCTP, dGTP, dTTP, 25 µM each; 75 µM ddGTP; 175 µM MgCl 2 ;<br />

pH 7.5).<br />

ddTTP (dATP, dCTP, dGTP, dTTP, 25 µM each; 1275 µM ddTTP; 1370 µM MgCl 2 ;<br />

pH 7.5).<br />

A second set of termination mixtures is shipped with the kit substituting 7-deaza-dGTP<br />

for dGTP.<br />

4. 0.5-mL thin-walled reaction tubes (N801-0537, Applied <strong>Bio</strong>systems, Foster City, CA).<br />

5. Eppendorf centrifuge.<br />

6. Thermal cycler for cycle sequencing (see Note 6 for reference).<br />

2.3. Preparation of Sequencing Gel<br />

1. 10× TBE: 0.9 M Tris, pH 8.3, 0.9 M boric acid, and 25 mM EDTA (6).<br />

2. 10% Ammonium persulfate (w/v) (should be prepared freshly before use).<br />

3. Sigmacote (SL-2, Sigma).<br />

4. GelMix 8 (#5545UA, Gibco BRL, Gaithersburg, MD), containing 7.6% acrylamide (w/v),<br />

0.4% N,N′-methylene bis-acrylamide, 7.0 M urea, 100 mM Tris-borate, pH 8.3, 1 mM<br />

Na 2 EDTA, 3 mM TEMED.<br />

5. 10-mL disposable syringes with needles.<br />

6. Scotch electrical tape 50 mm (FE-5000-0409-1, 3 M).<br />

7. Plastic foil (e.g., Saran Wrap).<br />

2.4. Gel Electrophoresis and Visualization of Sequencing Results<br />

1. 1× Tris-buffered saline (TBS) 50 mM Tris-HCl, pH 7.5, 150 mM NaCl.<br />

2. Alkaline phosphatase reaction buffer: 0.1 M Tris-HCl, pH 9.5, 50 mM MgCl 2 ,<br />

0.1 M NaCl.<br />

3. Stock solutions of 70% (v/v) and 100% N,N-dimethyl formamide.<br />

4. Digoxigenin Detection Kit for Glycoconjugate and Protein Analysis (#1210220, Boehringer<br />

Mannheim), containing: antidigoxigenin antibodies conjugated to alkaline phosphatase;<br />

blocking reagent (purified casein fraction); nitroblue tetrazoliumchloride, 77 mg/mL in<br />

70% N,N-dimethyl formamide; and X-Phosphate (5-bromo-4-chloro-3-indolylphosphate,<br />

4-toluidinium salt), 50 mg/mL in 100% N,N-dimethyl formamide.<br />

5. Nylon membrane, positively charged (#1417240, Boehringer Mannheim).<br />

6. Whatman chromatography paper (#3030917, 3MM Whatman International Ltd., Maidstone,<br />

UK).<br />

7. Plastic hybridization bags.<br />

8. Scalpel blade.<br />

9. Standard sequencing equipment: sequencing apparatus (Model S2, Gibco BRL) and<br />

high-voltage power supply (PS 9009, Gibco BRL).<br />

10. Ultraviolet transilluminator (302 nm) for DNA crosslinking (Hoefer).<br />

11. Eppendorf centrifuge.<br />

12. Heating block or thermal cycler for denaturation of DNA.<br />

13. Electric sealer for closing hybridization bags.


350 Kösel et al.<br />

3. Methods<br />

3.1. Purification of Sequencing Templates<br />

1. PCR is performed according to established protocols (7). After PCR amplification, PCR<br />

products are purified away from excess nucleotides and primers using either spin column<br />

chromatography (Wizard PCR Preps DNA Purification System, Promega, Madison, WI)<br />

or ultrafiltration (Microcon-30, Amicon). The purification systems are used according to<br />

manufacturer’s recommendation (see Note 3).<br />

2. Measure DNA concentration using the fluorometer and Hoechst dye No. 33258. The dye<br />

is dissolved in 1× TNE (prepare two stocks, 0.1 and 1 µg/mL, respectively). Calf thymus<br />

DNA (100 or 1000 ng/µL) is used as a standard for calibration (see Note 4).<br />

3. Purification results may be checked by electrophoresis of samples through a 2% highmelting-point<br />

agarose gel.<br />

3.2. Sequencing Reactions<br />

1. Sequencing reactions are set up in a total volume of 20 µL containing the follow: 13 µL of<br />

DNA template solution (2–4 pmol); dilute with sterile, double-distilled water, if necessary;<br />

3 µL of 5′-digoxigenin end-labeled primer (1 pmol/µL); 2 µL of 10× reaction buffer<br />

(250 mM Tris-HCl, pH 8.0, 50 mM MgCl 2 ); and 2 µL of Taq polymerase (3 U/µL).<br />

2. For each sequencing reaction, transfer 4 µL of the above mixture to four thin-walled PCR<br />

tubes, each containing 2 µL of the respective termination mixture (ddATP, ddCTP, ddTTP,<br />

and ddGTP; see Note 5).<br />

3. Overlay samples with 20 µL of mineral oil, and centrifuge for a few seconds in an<br />

Eppendorf tube, centrifuge at full speed.<br />

4. Cycle sequencing is performed using a thermal cycling protocol empirically optimized for<br />

the T m of the sequencing primer. When using sequencing primers of the same sequence<br />

and length as those used for PCR, a thermal cycling protocol identical to the one used for<br />

PCR usually gives good results (see Note 6).<br />

3.3. Preparation of Sequencing Gel<br />

For gel electrophoretic separation of sequencing reactions, an 8% denaturing<br />

polyacrylamide gel and standard sequencing equipment are used.<br />

1. Clean glass plates thoroughly with 70% ethanol. Cover inner surface of one plate with<br />

a few drops or Sigmacote and let evaporate for 5 to 10 min. Set up glass plates with<br />

spacers and seal edges airtight using Scotch electrical tape. Put two strong metal clamps<br />

on each side.<br />

2. Add 450 µL of 10% ammonium persulphate to one bottle (75 mL) of GelMix 8 and<br />

mix gently. Slowly fill the space between the glass plates (avoid air bubbles!). Let gel<br />

polymerize for at least 1 h at room temperature. (Wear protective gloves when handling<br />

unpolymerized acrylamide.)<br />

3. After polymerization of gel, remove clamps and electrical tape from the lower end of<br />

the glass plates and transfer gel to sequencing apparatus. Fill the upper and lower buffer<br />

chambers with 500 mL of 1× TBE each. Rinse sample wells of gel with 1× TBE using a<br />

10-mL disposable syringe with needle.<br />

3.4. Gel Electrophoresis and Visualization of Sequencing Results<br />

3.4.1. Electrophoresis and Contact Blotting<br />

1. Add 3 µL of loading buffer to each tube containing the sequencing reactions with the<br />

respective “A,” “C,” “G,” and “T” termination mixtures. Centrifuge for a few seconds


Digoxigenin 351<br />

to mix, and denature samples for at least 2 min at 85°C using a heating block or<br />

thermal cycler.<br />

2. Transfer 6 µL of each sample to the wells of the sequencing gel. Run gel at 35 mA for 2 to<br />

4 h (until the upper dye front has reached the bottom end of the gel).<br />

3. Remove glass plates from electrophoresis apparatus. Carefully take off the glass plate<br />

covered with Sigmacote on the inner side using a scalpel blade as a lever. Cover gel with<br />

nylon membrane (avoid air bubbles!) and cover with one layer of Whatman filter paper.<br />

Put on second glass plate and add approx 2 kg of weight. Leave for 30 min (see Note 7).<br />

Protect nylon membrane with one layer of plastic foil.<br />

4. Crosslink nylon membrane (plastic wrapped, blotted side down) on an ultraviolet light box<br />

at 302 nm for 3 min (see Note 8). The membrane may now be air-dried, sealed in a plastic<br />

bag (see Note 9), and stored at 4°C until further use.<br />

3.4.2. Visualization of Sequencing Bands<br />

Visualization of sequencing results is performed in plastic hybridization bags (one<br />

blot per bag) at room temperature using the Digoxigenin Detection Kit (Boehringer).<br />

After crosslinking, wet blots may be used immediately for immunological detection.<br />

The following recipe applies to a 4∞4 lane gel (gently agitate membrane, except during<br />

the final step when incubating with substrate solution).<br />

1. Incubate nylon membrane (in hybridization bag) for at least 30 min in approx 30 mL of<br />

blocking solution (0.5 g/100 mL TBS; see Note 10).<br />

2. Wash blot three times in approx 100 mL of TBS (10 min each).<br />

3. Incubate membrane with 50 mL of alkaline phosphatase-conjugated antidigoxigenin<br />

antibody for 1 h (50 µL antiserum/50 mL of TBS).<br />

4. Wash three times in TBS (20 min each).<br />

5. Prepare 30 mL of substrate solution (mix immediately before use): 150 µL of NBT<br />

(77 mg/mL in 70% N,N-dimethyl formamide [v/v]); 112 µL X-phosphate (50 mg/mL in<br />

N,N-dimethyl formamide); and 30 mL of alkaline phosphatase buffer, pH 9.5.<br />

Thoroughly remove washing solution from membrane. Add substrate solution. During<br />

this incubation step, cover plastic bag with black light protector and do not agitate (see<br />

Note 11).<br />

6. Stop developing process when faint sequencing bands are beginning to appear. Add approx<br />

1 L of deionized water to hybridization bag. Cut bag open and carefully remove membrane.<br />

Allow membrane to air-dry on sheets of Whatman filter paper or paper towels. Store<br />

stained, dried membranes in the dark (see Note 12). An example of a stained membrane<br />

is shown in Fig. 2.<br />

4. Notes<br />

1. Molecular biology-grade reagents are used for all experiments.<br />

2. Oligonucleotide primers for PCR and sequencing may be designed by hand, but use of a<br />

computer program, such as Oligo (National <strong>Bio</strong>sciences, Oslo, Norway), greatly facilitates<br />

this task. We have had good experience with primers ranging from 18 to 25 bp in length<br />

for both PCR and sequencing. The sequencing primer may be internally nested or identical<br />

in position and/or length to the primer used for PCR. Primers used in our laboratory were<br />

synthesized on an Applied <strong>Bio</strong>systems 394 DNA synthesizer at the Genzentrum of the<br />

University of Munich. Sequencing primers were custom-made and hapten-labeled using<br />

digoxigenin-3-O-methylcarbonylε-amino-caproic acid-N-hydroxy-succinimide ester and<br />

a 5′-oligopeptide linker (#1333054, Boehringer Mannheim).


352 Kösel et al.<br />

Fig. 2. Sequencing results obtained with primers ND3 5′-TCC CCA CCA TCA TAG CCA-3′<br />

and ND4 5′-GGG TTT TGC AGT CCT TAG-3′, which amplify a 302-bp segment of the<br />

mitochondrial ND2 gene (3). Sequencing results were visualized using alkaline phosphataseconjugated<br />

antibodies directed against digoxigenin and NBT/X-phosphate as an enzymatic<br />

substrate. Contact blot of an 8% standard sequencing gel.<br />

3. Purification of PCR products using the Wizard PCR Preps DNA Purification System: Add<br />

50 to 100 µL of PCR product to 100 µL of purification buffer in an Eppendorf tube and<br />

mix. Then, add 1 mL of purification resin and mix thoroughly. Pipet mixture into a 2-mL<br />

syringe and push into a Wizard Prep column. Wash column with 2 mL of 80% isopropanol<br />

and spin for 20 s in an Eppendorf centrifuge at full speed. Let isopropanol evaporate (takes<br />

approx 5 min), and transfer column to new Eppendorf tube. Elute DNA by adding 50 µL<br />

of 1× TE to the column. After 1 min, spin column for 20 s to recover DNA completely.<br />

When ultrafiltration is used for purification of PCR products, 50 to 100 µL of PCR<br />

product are diluted with approx 400 µL of double-distilled water and transferred to a<br />

Microcon-30 concentrator. Centrifugation of the Microcon-30 column is done according<br />

to manufacturer’s recommendation (e.g., 10 min at 12,000g) in an Eppendorf centrifuge.<br />

Important: Avoid contamination of both types of columns with mineral oil used for<br />

overlaying PCR.<br />

4. A fluorometer is used to measure nanogram amounts of DNA. The TKO 100 (Hoefer) is<br />

a useful and reasonably priced alternative to full-sized spectrophotometers for measuring<br />

concentrations of double-stranded DNA. DNA concentrations as low as 10 ng/µL may be<br />

reliably determined using a standard setup of the TKO. For measuring the concentration of<br />

PCR products, the fluorometer is first calibrated with DNA standard (100 or 1000 ng/µL<br />

calf thymus DNA, depending on the DNA concentrations to be determined): 1 µL of


Digoxigenin 353<br />

standard DNA is mixed with 1 mL of 1× TNE containing Hoechst dye No. 33258 and<br />

added to the quartz cuvet. Concentration of the Hoechst dye is 0.1 or 1 µL/mL in 1X TNE,<br />

depending on the concentration of the calibration standard.<br />

5. Termination mixtures containing 7-deaza-dGTP are used to avoid band compression<br />

artifacts when sequencing guanine–cytosine-rich regions.<br />

6. Similar to PCR, the success of Taq cycle sequencing depends on buffer conditions,<br />

especially the concentration of Mg 2+ ions and pH. Buffer conditions for PCR and cycle<br />

sequencing may be optimized using commercial “optimizer” kits (e.g., K1220-01,<br />

Invitrogen, Leek, Netherlands). We have used Taq polymerases from Perkin–Elmer and<br />

Boehringer Mannheim with comparable success for both PCR and sequencing. Thermal<br />

cyclers from Perkin–Elmer (TC 480) and <strong>Bio</strong>metra (Trio Thermo block) also gave<br />

comparable results in our hands. Sequencing protocols using smaller numbers of cycles,<br />

e.g., 20 to 25 may reduce the intensity of shadow bands.<br />

7. Do not extend blotting time and do not use higher weights.<br />

8. Protect eyes and skin against ultraviolet light (wear goggles, mask, coat, and gloves)!<br />

9. To save on reagents, hybridization bags should be tightly sealed. However, leave one “long<br />

end,” because bags will need to be reopened and resealed during subsequent incubation<br />

steps.<br />

10. Blocking reagent (Boehringer Mannheim) should be prepared freshly before use. Dissolve<br />

blocking reagent at 50°C in TBS, and let cool down to room temperature. Blots can be<br />

stored in blocking solution at 4°C for a few days.<br />

11. To avoid diffuse or spotty background, the volume of the substrate solution should be<br />

sufficiently large to cover the nylon membrane completely. Also, avoid folds in the<br />

hybridization bag. Sequencing bands should become visible within 15 min on incubation.<br />

Do not move the membrane during that time. Because the visualization process involves an<br />

enzymatic reaction, lowering the temperature of the incubation solutions may lengthen incubation<br />

times. We recommend developing at room temperature and checking the intensity of<br />

the stained sequencing bands by briefly lifting the dark cover from time to time.<br />

12. The intensity of the sequencing bands may increase during the first few hours after the final<br />

incubation has been finished (if alkaline phosphatase and NBT/X-phosphate are used for<br />

visualization as in the present protocol). Therefore, the incubation with substrate solution<br />

should be stopped as soon as the first sequencing bands are clearly readable. Important:<br />

Stained, dried nylon membranes should not be exposed to sunlight because they will<br />

bleach rapidly. However, stained membranes can be stored safely for many months in<br />

the dark. For long-term documentation, we recommend taking photographs. Sequencing<br />

results may also be documented permanently by photocopying stained blots onto paper<br />

or overhead transparencies.<br />

References<br />

1. Kösel, S. and Graeber, M. B. (1993) Non-radioactive direct sequencing of PCR products<br />

amplified from neuropathological specimens. Brain Pathol. 3, 421– 424.<br />

2. Kösel, S. and Graeber, M. B. (1994) Use of neuropathological tissue for molecular<br />

genetic studies: parameters affecting DNA extraction and polymerase chain reaction. Acta<br />

Neuropathol. 88, 19–25.<br />

3. Kösel, S., Egensperger, R., Mehraein, P., and Graeber, M. B. (1994) No association of<br />

mutations at nucleotide 5460 of mitochondrial NADH dehydrogenase with Alzheimer’s<br />

disease. <strong>Bio</strong>chem. <strong>Bio</strong>phys. Res. Commun. 203, 745–749.<br />

4. Krishnan, B. R., Blakesley, R. W., and Berg, D. E. (1991) Linear amplification DNA<br />

sequencing directly from single phage plaques and bacterial colonies. Nucleic Acids Res.<br />

19, 1153.


354 Kösel et al.<br />

5. Kwok, S. and Higuchi, R. (1989) Avoiding false positives with PCR. Nature 339,<br />

237–238.<br />

6. Sambrook, J., Fritsch, E. F., and Maniatis, T. (1989) Molecular Cloning: A Laboratory<br />

Manual, 2nd ed., Cold Spring Harbor Laboratory Press, Cold Spring Harbor, New York.<br />

7. Saiki, R. K., Gelfand, D. H., Stoffel, S., Scharf, S. J., Higuchi, R., Horn, G. T., et al. (1988)<br />

Primer-directed enzymatic amplification of DNA with a thermostable DNA polymerase.<br />

Science 239, 487– 491.


Thermal Asymmetric PCR 355<br />

52<br />

Direct Sequencing by Thermal Asymmetric PCR<br />

Georges-Raoul Mazars and Charles Theillet<br />

1. Introduction<br />

Direct sequencing of polymerase chain reaction (PCR) products (1) has proven to<br />

be a powerful method in the generation of nucleic acid sequence data. Using these<br />

techniques, it is possible to produce microgram quantities of pure target DNA and<br />

subsequently its nucleotide sequence in a few hours, even, theoretically, from one<br />

single RNA or DNA molecule. However, problems have been encountered, and these<br />

have been attributed to the strong tendency of the short double-strand DNA templates<br />

to reanneal. In fact, compared with double-stranded plasmid DNA, which can be<br />

permanently denatured by alkali treatment and then form intermolecular interactions<br />

compatible with good sequencing efficiency, optimized conditions for direct sequencing<br />

are required before reannealing with short PCR product.<br />

To obviate this, strategies have been developed, such as the generation of singlestrand<br />

DNA template by asymmetric PCR (2,3). Methods use either a disequilibrated<br />

concentration ratio between the two primers or a two-step amplification, both of which<br />

have their shortfalls. The first method is based on a large number of cycles, which is<br />

a potential source of misincorporation of errors, and optimized conditions enough to<br />

produce single-strand DNA that are strongly primer dependent. Moreover, it often has<br />

been the case that only one strand can easily be sequenced. The second case requires<br />

two physical separation steps, in which product contamination may occur.<br />

Here, we propose a method combining the advantages of both symmetric and<br />

asymmetric PCR. It is based on a thermal asymmetry between the Tm of both primers.<br />

Annealing temperature of each primer is calculated with the formula: 69.3 + 0.41<br />

(%GC) – 650/L, with L = primer length. PCR primers are designed to obtain a difference<br />

in T m of at least 10°C. In the first step, double-stranded material is produced<br />

during 20 to 25 cycles (to minimize the yield of spurious products) using the lower<br />

Tm. During the second step, single-stranded DNA is generated using the higher<br />

T m (Fig. 1).<br />

Consequently, one primer is dropped out and linear amplification is obtained.<br />

The final quantity of single-stranded product is comparable with the one produced<br />

by Gyllensten and Erlich’s method (2). We applied thermal asymmetry to several<br />

sequences, which, in our hands, were difficult to sequence both from double-stranded<br />

DNA or with the Gyllensten and Erlich asymmetric PCR products (Fig. 2).<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

355


356 Mazars and Theillet<br />

Fig. 1. Schematic representation of thermal asymmetric PCR.<br />

These PCR fragments comprised the following: (1) exons 7 and 8 of the p53 gene<br />

and (2) exon 1 of HRAS. This latter sequence is particularly GC-rich, and as part<br />

of another study, primer A was synthesized with a 40-bp GC stretch (to make a GC<br />

clamp). In conclusion, thermal asymmetric PCR allows direct sequencing of both<br />

strands with high reproducibility and reduced risk of contamination.<br />

2. Materials<br />

All solutions should be made according to the standard required for molecular<br />

biology, such as molecular biology-grade reagents and sterile distilled water. All<br />

reagents for sequencing are available commercially.<br />

2.1. PCR Amplification<br />

1. Primers were synthesized on Applied Amplifications, and PCR was performed on a<br />

Perkin–Elmer Cetus thermal cycler.<br />

2. 1× Taq polymerase buffer: 10 mM Tris-HCl, pH 8.3, 2 mM MgCl 2 , 50 mM KCl, 0.01%<br />

gelatin; 100 µM of dNTPs.<br />

3. Taq polymerase was purchased from Perkin–Elmer and used at 1 U/reaction.


Thermal Asymmetric PCR 357<br />

Fig. 2. Autoradiograph of a PCR product sequenced by thermal asymmetric PCR.<br />

2.2. Purification and Sequencing of the PCR Product<br />

2.2.1. Sequencing Reagents<br />

1. Annealing buffer (5× concentrate): 200 mM Tris-HCl, pH 7.5, 100 mM MgCl 2 , 250 mM<br />

NaCl, 0.1 M dithiothreitol (DTT).<br />

2. Labeling nucleotide mixture (one for each dideoxy nucleotide): Each mixture contains<br />

80 µM dGTP, 80 µM dATP, 80 µM dTTP, 80 µM dCTP, and 50 mM NaCl. In addition,<br />

the “G” mixture contains 8 µM dideoxy-dGTP; the “A” mixture, 8 µM ddATP; the “T,”<br />

8 µM ddTTP; and the “C” 8 µM ddCTP.<br />

3. Stop solution: 95% formamide, 20 mM EDTA, 0.05% bromophenol blue, and 0.05%<br />

xylene cyanol FF.<br />

4. Labeled dATP is [α- 35 S] dATP from Amersham, and specific activity should be 1000<br />

to 1500 Ci/mol.<br />

2.2.2. Purification<br />

Purification should be performed in Centricon 30 column (Amicon).<br />

3. Methods<br />

3.1. PCR Conditions<br />

PCR conditions should be as follows: in a total volume of 25 µL, 20 pmol of each<br />

primer, 1× Taq polymerase buffer, and 1 U of Taq polymerase were incubated with 50 ng


358 Mazars and Theillet<br />

of genomic DNA. P53 primer A1 cttagtacctgaagggtgaaatattc (T m 1 = 60°C), P53 primer<br />

B1 gtagtggtaatctactgggacggaacagc (T m 2 = 69°C), P53 primer A2 taatctactgggacgga<br />

(T m 3 = 50°C), P53 primer B2 cccaagacttagtacctgaagggtg (T m 4 = 64°C). Cycling<br />

conditions were for 25 cycles: 92°C (30 s); T m 1 or 3 (30 s); 72°C (90 s), followed by<br />

10 cycles: 92°C (30 s); T m 2 or 4 (30 s); 72°C (90 s).<br />

3.2. Preparation of the PCR Product<br />

1. Transfer PCR product directly by pipetting in a Centricon 30 and add 2 mL of water.<br />

2. Spin at 5000g in a fixed-angled rotor in a Beckman-type centrifuge for 30 min at room<br />

temperature.<br />

3. Add 2 mL of water. Spin again for 30 min and invert column and spin for 5 min at 1500g.<br />

This procedure efficiently removes the excess of dNTPs from the PCR. Volume recovered<br />

is typically 20 to 50 µL.<br />

4. Typically, 7 µL of this purified product are used for single-strand sequencing according to<br />

the manufacturer’s directions of United States <strong>Bio</strong>chemicals (Cleveland, OH).<br />

3.3. Sequencing Protocol<br />

3.3.1. Annealing Template and Primer<br />

1. For each template, a single annealing (and subsequent labeling) reaction is used. Combine<br />

the following:<br />

a. Primer 0.5 pmol (1 µL)<br />

b. DNA 7 µL<br />

c. Annealing buffer 2 µL<br />

2. Warm the capped tube to 65°C for 2 min, and then allow the mixture to cool slowly to<br />

room temperature over a period of about 30 min.<br />

3.3.2. Labeling Reaction<br />

1. To the annealed template-primer add the following:<br />

a. DTT (0.1M) 1 µL<br />

b. Labeling nucleotide mix 2 µL<br />

c. [α-α- 35 S] dATP 5 µCi (typically 0.5 µL)<br />

d. Sequenase 3 U from United States <strong>Bio</strong>chemicals<br />

2. Total volume should be approx 15 µL; mix thoroughly and incubate for 2 to 5 min at<br />

room temperature.<br />

3.3.3. Termination Reactions<br />

1. Label four tubes “A,” “C,” “G,” and “T.” Fill each with 2.5 µL of the appropriate dideoxy<br />

termination mixture.<br />

2. When the labeling reaction is complete, transfer 3.5 µL of it to the tube (prewarmed to<br />

37°C) labeled G. Similarly, transfer 3.5 µL of the labeling reaction to each of the other<br />

three tubes (A, T, and C).<br />

3. After 2 to 5 min of incubation at 37°C, add 4 µL of stop solution to each termination<br />

reaction, mix, and store on ice.<br />

4. To load the gel, heat the samples to 75 to 80°C for 2 min, and load 2 to 3 µL in each<br />

lane. Prerun a sequencing gel for 30 min, load, and run until bromophenol is just out<br />

of the gel.<br />

5. Fix gel as usual and dry on Whatman 3MM paper. Correct sequencing yields a detectable<br />

signal using a bench-top Geiger counter.<br />

6. Expose overnight without Saran paper at room temperature.


Thermal Asymmetric PCR 359<br />

4. Notes<br />

1. Instability of diluted solution of primers conserved at –20°C can sometimes be problematic.<br />

We recommend storing oligonucleotides as a dried powder and resuspending them in<br />

water prior to use.<br />

2. Estimation of the yield of single-strand DNA produced can be achieved by Southern<br />

blotting: run a 2% agarose gel, blot following standard conditions, and probe with one of<br />

the PCR primers. Two bands should appear if you use higher T m primer or only one band<br />

if you use lower T m primer (corresponding to double-stranded DNA). An alternative<br />

strategy is to add α-dCTP[ 32 P] for the second step of amplification at high temperature:<br />

labeled single-stranded DNA should be exclusively produced.<br />

3. We also sequenced PCR fragments after SSCP analysis. In this case, shifted SSCP bands<br />

were excised from the gel with a sterile razor blade and eluted in 50 µL of distilled water<br />

for 1 h at 65°C. A 1.5-µL aliquot of the eluate was subjected to thermal asymmetric PCR.<br />

4. A good sequence can be obtained even if no primer is added for the sequencing reaction.<br />

The reason is that the low T m primer from PCR is not completely removed by Centricon<br />

30 purification.<br />

5. The present protocol has been optimized for classical radioactivity labeled DNA sequencing,<br />

but should easily be adapted to automated fluorescent sequencing using fluorescent<br />

dye terminators.<br />

References<br />

1. Saiki, R. K., Gelfand, D. H., Stoffel, S., Scharf, S. F., Higuchi, R., Horn, R. T., et al. (1988)<br />

Primer-directed enzymatic amplification of DNA with a thermostable DNA polymerase.<br />

Science 239, 487– 491.<br />

2. Gyllensten, U. B. and Erlich, H. A. (1988) Generation of single-stranded DNA by the<br />

polymerase chain reaction and its application to direct sequencing of the HLA-DQA locus.<br />

Proc. Natl. Acad. Sci. USA 85, 7652–7656.<br />

3. Wilson, R. K., Chen, C., and Hood, L. (1990) Optimization of asymmetric polymerase<br />

chain reaction for rapid fluorescent DNA sequencing. <strong>Bio</strong>Techniques 8, 184–189.


360 Mazars and Theillet


Solid-Phase Minisequencing 361<br />

53<br />

Analysis of Nucleotide Sequence Variations<br />

by Solid-Phase Minisequencing<br />

Anu Suomalainen and Ann-Christine Syvänen<br />

1. Introduction<br />

The Sanger dideoxynucleotide sequencing method has been simplified by a number<br />

of methodological improvements, such as the use of the polymerase chain reaction<br />

(PCR) technique for generating DNA templates in sufficient quantities followed<br />

by affinity-capture techniques for convenient and efficient purification of the PCR<br />

fragments for sequencing, or altenatively the use of cyclic Sanger sequencing reactions<br />

that are easy to automate with laboratory robots, and the development of instruments<br />

for automatic on-line analysis of fluorescent products of the sequencing reactions<br />

(references to this book, solid phase sequencing, cyclic sequencing). Despite these<br />

technical improvements, the requirement for gel electrophoretic separation remains<br />

an obstacle when sequence analysis of large numbers of samples are needed, as in<br />

DNA diagnosis, or in the analysis of sequence variation for genetic, evolutionary, or<br />

epidemiological studies.<br />

We have developed a method for analyzing DNA fragments that differ from each<br />

other in one or a few nucleotide positions (1) denoted solid-phase minisequencing, in<br />

which gel electrophoretic separation is avoided. Analogous to the methods for solidphase<br />

sequencing of PCR products, the solid-phase minisequencing method is based<br />

on PCR amplification using one biotinylated and one unbiotinylated primer, followed<br />

by affinity-capture of the biotinylated PCR product on an avidin- or streptavidincoated<br />

solid support. The nucleotide at the variable site is detected in the immobilized<br />

DNA fragment by a primer extension reaction: A detection step primer that anneals<br />

immediately adjacent to the nucleotide to be analyzed is extended by a DNA polymerase<br />

with a single labeled nucleotide complementary to the nucleotide at the variable site<br />

(Fig. 1). The amount of the incorporated label is measured, and it serves as a specific<br />

indicator of the nucleotide present at the variable site.<br />

We have used the solid-phase minisequencing method for detecting numerous<br />

mutations causing human genetic disorders, for analyzing allelic variation in genetic<br />

linkage studies, and for identifying individuals (2–4). The protocol described below<br />

is generally applicable for detecting any variable nucleotide. The method suits well for<br />

analyzing large numbers of samples because it comprises simple manipulations in<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

361


362 Suomalainen and Syvänen<br />

Fig. 1. Steps of the solid-phase minisequencing method. 1. PCR with one biotinylated (black<br />

ball) and one unbiotinylated primer. 2. Affinity-capture of the biotinylated PCR product in<br />

streptavidin-coated microtiter wells. 3. Washing and denaturation. 4. The minisequencing primer<br />

extension reaction. 5. Measurement of the incorporated label. 6. Calculation of the result.


Solid-Phase Minisequencing 363<br />

a microtiter plate or test tube format and the result of the assay is obtained as an<br />

objective numeric value, which is easy to interpret. Furthermore, the solid-phase<br />

minisequencing method allows quantitative detection of a sequence variant present as a<br />

minority of less than 1% in a sample (2,3,5,6). We have used the sensitive quantitative<br />

analysis for detecting point mutations in malignant cells present as a minority in a cell<br />

population (5) and for analyzing heteroplasmic mutations of mitochondrial DNA (3,6).<br />

The high sensitivity is an advantage of the minisequencing method compared with<br />

dideoxynucleotide sequencing, in which a sequence variant must be present as 10 to<br />

20% of a mixed sample to be detectable. A limitation of the solid-phase minisequencing<br />

method is that it is restricted to analyzing variable nucleotides only at positions<br />

predefined by the detection step primers used. The method is based on the use of<br />

equipment and reagents that are available from common suppliers of molecular<br />

biological products, facilitating easy setup. In the future, high-throughput analysis of<br />

nucleotide sequence variation will be performed by rapid, automatic methods based<br />

on homogeneous detection principles or alternatively using methods in microarray<br />

or chip formats. The minisequencing reaction principle is applicable for both types<br />

of assay formats (7,8).<br />

2. Materials<br />

2.1. Equipment<br />

1. Programmable heat block, and facilities to avoid contamination in PCR.<br />

2. Microtiter plates with streptavidin-coated wells (e.g., Combiplate 8, Labsystems, Helsinki,<br />

Finland; see Note 1).<br />

3. Multichannel pipet and microtiter plate washer (optional).<br />

4. Shaker at 37°C.<br />

5. Water bath or incubator at 50°C.<br />

6. Liquid scintillation counter.<br />

2.2. Reagents<br />

All the reagents should be of standard molecular biology grade. Use sterile distilled<br />

or deionized water.<br />

1. Thermostable DNA polymerase. We use Thermus aquaticus (5 U/µL, Promega or Perkin–<br />

Elmer-ABI) or Thermus brockianus (Dynazyme II, 2 U/µL, Finnzymes, Espoo, Finland)<br />

DNA polymerase. Store at –20°C (see Note 2).<br />

2. 10× concentrated DNA polymerase buffer: 500 mM mM MgCl 2 , 1% (v/v) Triton X-100,<br />

and 0.1% (w/v) gelatin or 10× concentrated buffer supplied with the DNA polymerase<br />

enzyme. Store at –20°C.<br />

3. dNTP mixture: 2 mM dATP, 2 mM dCTP, 2 mM dGTP, and 2 mM dTTP stored at –20°C.<br />

4. PBS/Tween: 20 mM sodium phosphate buffer, pH 7.5, and 0.1% (v/v) Tween 20 store at<br />

4°C. 50 mL is enough for several full-plate analyses.<br />

5. TENT (washing solution): 40 mM Tris-HCl, pH 8.8, 1 mM EDTA, 50 mM NaCl, and<br />

0.1% (v/v) Tween 20. Store at 4°C. Prepare 1 to 2 L at a time, which is enough for several<br />

full-plate analyses.<br />

6. NaOH (50 mM; make fresh every 4 wk) stored in a plastic vial at room temperature<br />

~20°C).<br />

7. [ 3 H]-labeled deoxynucleotides (dNTPs): dATP to detect a T at the variant site, dCTP to


364 Suomalainen and Syvänen<br />

detect a G, etc. (Amersham-Pharmacia <strong>Bio</strong>tech; [ 3 H]dATP, TRK 633; dCTP, TRK 625;<br />

dGTP, TRK 627; dTTP, TRK 576), stored at –20°C (see Note 3).<br />

8. Scintillation fluid (for example Hi-Safe II, Wallac, Turku, Finland) stored at room<br />

temperature (~20°C).<br />

2.3. Primer Design<br />

1. PCR primers: One PCR primer of each pair is biotinylated at its 5′ end during the synthesis<br />

using a biotin-phosphoramidite reagent (for example Amersham–Pharmacia <strong>Bio</strong>tech or<br />

Perkin-Elmer–ABI; see Note 4).<br />

2. The detection step primer for the minisequencing analysis is a 20-mer oligonucleotide<br />

complementary to the biotinylated strand of the PCR product, designed to hybridize with<br />

the 3′ end immediately adjacent to the variant nucleotide to be detected (see Fig. 1)<br />

The minisequencing primer should be at least five nucleotides nested in relation to the<br />

unbiotinylated PCR primer.<br />

3. Methods<br />

3.1. PCR for Solid-Phase Minisequencing Analysis<br />

The PCR is performed according to routine protocols, except that the amount of<br />

the biotin-labeled primer used is reduced not to exceed the biotin-binding capacity of<br />

the microtiter well (see Note 1). For a 50-µl PCR, we used 10 pmol of biotin-labeled<br />

primer and 50 pmol of the unbiotinylated primer. The PCR should be optimized (i.e.,<br />

the annealing temperature and template amount) to be efficient and specific. To be<br />

able to use [ 3 H] dNTPs, which are low-energy β-emitters, for the minisequencing<br />

analysis, 1/10 of the PCR product should produce a single visible band after agarose<br />

gel electrophoresis, stained with ethidium bromide. There is no need for purification<br />

of the PCR product before the minisequencing analysis.<br />

3.2. Solid-Phase Minisequencing Analysis<br />

1. Affinity capture: Transfer 10-µL aliquots of the PCR product and 40 µL of the PBS/Tween<br />

solution to two streptavidin-coated microtiter wells (see Note 5). Include a control reaction,<br />

that is, a well with no PCR product. Seal the wells with a sticker and incubate the plate<br />

at 37°C for 1.5 h with gentle shaking.<br />

2. Discard the liquid from the wells and tap the wells dry against a tissue paper.<br />

3. Wash the wells three times at room temperature as follows: pipet 200 µL of TENT solution<br />

to each well, discard the washing solution and empty the wells thoroughly between the<br />

washings (see Note 6).<br />

4. Denature the captured PCR product by adding 100 µL of 50 mM NaOH to each well,<br />

incubate at room temperature for 3 min. Discard the NaOH and wash the wells as in<br />

step 3 above.<br />

5. Prepare for each DNA fragment to be analyzed two 50-µL mixtures of nucleotide-specific<br />

minisequencing solution, one for detection of the normal and one for the mutant nucleotide<br />

(see Note 7). Mix 5 µL of 10× Taq DNA polymerase buffer, 10 pmol of detection step<br />

primer, 0.2 µCi (usually equals to 0.2 µL) of one [ 3 H] dNTP, 0.1 U of Taq DNA polymerase,<br />

and dH 2 O to a total volume of 50 µL. It is obviously convenient to prepare master mixes<br />

for the desired number of analyses with each nucleotide.<br />

6. Pipet 50 µL of one nucleotide-specific mixture per well, incubate the plate at 50°C for<br />

10 min in a water bath or 20 min in an oven (see Note 8).<br />

7. Discard the contents of the wells and wash them as in step 3.


Solid-Phase Minisequencing 365<br />

8. Release the detection step primer from the template by adding 60 µL of 50 mM NaOH and<br />

incubating for 3 min at room temperature.<br />

9. Transfer the eluted primer to the scintillation vials, add scintillation reagent, and measure<br />

the radioactivity, that is, the amount of incorporated label, in a liquid scintillation counter<br />

(see Note 9).<br />

10. The result is obtained as counts per minute (cpm) values. The cpm value of each reaction<br />

expresses the amount of the incorporated [ 3 H] dNTP. Calculate the ratio (R) between the<br />

mutant and normal nucleotide cpms. In a sample of a subject homozygous for the mutant<br />

nucleotide the R will be >10, in a homozygote for the normal nucleotide R


366 Suomalainen and Syvänen<br />

temperatures for the primer annealing may be required.<br />

9. Streptavidin-coated microtiter plates made of scintillating polystyrene are available (Wallac,<br />

Finland). In this case the final washing, denaturation and transfer of the eluted detection<br />

primer can be omitted, but a scintillation counter for microtiter plates is needed (11).<br />

10. The ratio between the cpm values for the two nucleotides reflects the ratio between the<br />

two sequences in the original sample. Therefore, the solid-phase minisequencing method<br />

can be used for quantitative PCR analyses (2–6). The R value is affected by the specific<br />

activities of the [ 3 H] dNTPs used, and if either the mutant or the normal sequence allows<br />

the detection step primer to be extended by more than one [ 3 H] dNTP, this will obviously<br />

also affect the R value. Both of these factors can easily be corrected for when calculating<br />

the ratio between the two sequences. Alternatively, a standard curve is constructed by<br />

mixing the two sequences in known ratios and plotting the obtained R-values as a function<br />

of the ratios to obtain a linear standard curve (3,5,6). The test results can then be interpreted<br />

from the standard curve without the need of taking the specific activities of the number of<br />

[ 3 H] dNTPs incorporated into account.<br />

References<br />

1. Syvänen, A.-C., Aalto-Setälä, K., Harju, L., Kontula, K., and Söderlund, H. (1990) A<br />

primer-guided nucleotide incorporation assay in the genotyping of apolipoprotein E.<br />

Genomics 8, 684–692.<br />

2. Suomalainen, A. and Syvänen, A.-C. (2000) Quantitative analysis of human DNA sequences<br />

by PCR and solid-phase minisequencing. Mol. <strong>Bio</strong>technol. 15, 123–131.<br />

3. Suomalainen, A., Kollmann, P., Octave, J.-N., Söderlund, H., and Syvänen, A.-C. (1993)<br />

Quantification of mitochondrial DNA carrying the tRNA<br />

Lys 8344 point mutation in myoclonus<br />

epilepsy and ragged-red-fiber disease. Eur. J. Hum. Genet. 1, 88–95.<br />

4. Syvänen, A.-C. (1999) From gels to chips: “Minisequencing” primer extension for analysis<br />

of point mutations and single nucleotide polymorphisms. Hum. Mutat. 13, 1–10.<br />

5. Syvänen, A-C., Söderlund, H., Laaksonen, E., Bengtström, M., Turunen, M., and Palotie, A.<br />

(1992) N-ras gene mutations in acute myeloid leukemia: accurate detection by solid-phase<br />

minisequencing. Int. J. Cancer 50, 713–718.<br />

6. Suomalainen, A., Majander, A., Pihko, H., Peltonen, L., and Syvänen, A.-C. (1993)<br />

Quantification of tRNA<br />

Leu 3243 point mutation of mitochondrial DNA in MELAS patients<br />

and its effects on mitochondrial transcription. Hum. Mol. Genet. 2, 525–534.<br />

7. Chen, X. and Kwok, P. Y. (1997) Template-directed dye-terminator incorporation (TDI)<br />

assay: a homogeneous DNA diagnostic method based on fluorescence resonance energy<br />

transfer. Nucleic Acids Res. 15, 347–353.<br />

8. Pastinen, T., Perola, M., Niini, P., Terwilliger, J., Salomaa, V., Vartiainen, E., et al. (1998)<br />

Array-based multiplex analysis of candidate genes reveals two independent and additive<br />

risk factors for myocardial infarction in the Finnish population. Hum. Mol. Genet. 7,<br />

1453–1462.<br />

9. Pastinen, T., Partanen, J., and Syvänen, A.-C. (1996) Multiplex, fluorescent solid-phase<br />

minisequencing for efficient screening of DNA sequence variation. Clin. Chem. 42,<br />

1391–1397.<br />

10. Sitbon, G., Hurtig, M., Palotie, A., Lönngren, J., and Syvänen, A.-C. (1997) A colorimetric<br />

mini-sequencing assay for the mutation in codon 506 of the coagulation factor V gene.<br />

Thrombosis Haemostasis 77, 701–703.<br />

11. Ihalainen, J., Siitari, H., Laine, S., Syvänen, A.-C., and Palotie, A. (1994) Towards automatic<br />

detection of point mutations: use of scintillating microplates in solid-phase minisequencing.<br />

<strong>Bio</strong>Techniques 16, 938–943.


Inosine-Containing Primers 367<br />

54<br />

Direct Sequencing with Highly Degenerate<br />

and Inosine-Containing Primers<br />

Zhiyuan Shen, Jingmei Liu, Robert L. Wells, and Mortimer M. Elkind<br />

1. Introduction<br />

Among the many techniques of cloning new genes, one approach involves degenerate<br />

primers (1–7). The approach usually requires the following three steps:<br />

1. Using degenerate primers to amplify part of the gene of interest by PCR: The degenerate<br />

primers’ sequences may be designed from known protein sequences or conserved regions<br />

of a gene family (e.g., ref. 2,4). Because deoxyinosine can base pair with all of the four<br />

deoxyribonucleotides, it has been substituted for specific nucleic acids in degenerate<br />

primers to reduce the number of different primer sequences that would otherwise be<br />

needed in the reaction (2,7,8).<br />

2. A determination of which amplified PCR product(s) is from the gene of interest: If the<br />

target gene and the primers are only partially homologous, a moderate annealing stringency<br />

in PCR is usually necessary to obtain amplification. Moderate stringency may result in<br />

multiple PCR products. Although from the size of the PCR products it may be possible<br />

to predict which is from the gene of interest, sequencing analysis of the PCR products<br />

may be required.<br />

3. The screening of a cDNA library using the correct PCR product as a probe and cloning<br />

the gene of interest.<br />

Sequencing the amplified PCR product is one of the most important steps in this<br />

approach to gene cloning. To sequence the PCR fragment amplified by degenerate<br />

inosine-containing primers, the PCR fragment may be cloned into a sequencing vector,<br />

such as M13 bacteriophage. Sequencing is straightforward if primers specific to the<br />

vector are used. Theoretically, this method allows any unknown cloned DNA fragment<br />

to be sequenced. However, the Taq polymerase, which is used to amplify the target<br />

fragment, is thought to have relatively high misincorporation rates for dNTPs, or ~10 –4 .<br />

Hence, it is possible that a copy of the product may contain one or more incorrect<br />

nucleotides. If such a copy has been cloned into the sequencing vector, the resulting<br />

sequencing data would be incorrect for that particular clone. Direct sequencing of PCR<br />

products can circumvent this problem because most of the fragments are exact replicas<br />

of the target molecule. Thus, the majority of the products used for sequencing would<br />

have the right nucleotide at a specified position and result in the correct sequencing<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

367


368 Shen et al.<br />

ladder. Also, direct sequencing of PCR products avoids the time-consuming cloning<br />

of PCR products, and most of the available direct PCR sequencing protocols require<br />

relatively small amounts of template.<br />

Many protocols are available for the direct sequencing of PCR products. Most of<br />

the protocols require specific sequencing primers. However, this means at least part of<br />

the specific base sequence of the template is needed. This requirement may not be met<br />

and, in many cases, <strong>info</strong>rmation about the internal sequence of a gene may be lacking.<br />

This shortcoming may apply to the PCR products of degenerate inosine-containing<br />

primers of the cDNA of a new gene. Therefore, one may be forced to use the same<br />

degenerate inosine-containing primers that were used in the PCR step for direct<br />

sequencing. When primers have low degeneracy, they may be treated as sequencespecific<br />

primers, and some of the direct-sequencing protocols, such as those described<br />

in this volume, may be used with success. When only highly degenerate inosinecontaining<br />

primers are available, these methods may not succeed.<br />

To sequence a PCR product amplified via the use of a highly degenerate inosinecontaining<br />

primers, several general factors must be kept in mind.<br />

1. The sequencing primer(s) must anneal specifically to one site on the DNA fragment that is<br />

to be sequenced, that is, a secondary annealing site must be avoided. Therefore, stringent<br />

primer annealing temperatures are necessary.<br />

2. A sufficient quantity of the specific primer should anneal to the correct site. Consequently,<br />

the primer annealing temperature cannot be too high.<br />

3. Reassociation of the double-stranded DNA template should be minimized. This requirement<br />

generally can be met by using optimal PCR protocols for the sequencing reactions.<br />

To perform the requirements above, a primer-labeling method, in which the primer<br />

is labeled at the 5′ end, may be worth considering. Linear PCR is used to generate<br />

the labeled dideoxynucleotide-terminated sequences (9,10). The use of this method<br />

minimizes problems of template reassociation and/or mismatching of the primer<br />

because the annealing time is relatively short. Also, the annealing temperature is higher<br />

than it would be in most protocols that use DNA polymerases other than Taq, such as<br />

T4 DNA polymerase, but the method requires a 5′ end-labeling step for which 35 S is<br />

generally not suitable compared with 32 P because of its lower specific activity and the<br />

lesser efficiency with which some enzymes label 5′ ends with α- 35 S ATP versus α- 32 P<br />

ATP. Only 32 P can be used, even though its greater radiation hazard owing to its higher<br />

β-particle emission and its shorter half-life make it less convenient. Furthermore,<br />

when highly degenerate primers are used, higher primer concentrations in the reaction<br />

mixture are needed to insure that sufficient specific priming will occur. The preceding<br />

increases the hazard as well as the cost.<br />

To assure that sequencing primer(s) anneal to a DNA template specifically, to<br />

eliminate the need for 5′ end labeling, and to avoid reassociation of the double-stranded<br />

DNA template, a two-step cycle-sequencing protocol is described to sequence products<br />

amplified with degenerate inosine-containing primers. This method uses the same<br />

degenerate primers that were used in PCR amplification. The method can be broken<br />

down into two steps of linear PCR. The first step is for labeling the primers, and the<br />

second is for the random dideoxy termination. As shown in Fig. 1, in the first step,<br />

primers were extended and labeled with α- 35 S -dATP. The extension is limited and<br />

performed under conditions of high stringency, low dNTP concentration, and a short


Inosine-Containing Primers 369<br />

Fig. 1. Procedure of cycle sequencing with degenerate inosine-containing primers. Two<br />

linear PCR steps are involved. 1. Label PCR uses low dNTP concentrations, a low temperature,<br />

and short times for primer annealing/elongation to produce incomplete extension of specific<br />

primers. As a result, specific primers are labeled and extended. The extended and labeled<br />

primers have a higher melting temperature than the native printers. 2. Termination PCR using<br />

a higher annealing/elongation temperature and is performed with higher dNTP concentrations<br />

and in the presence of ddNTPs. Only the extended and labeled primers are involved in the<br />

termination reaction.<br />

interval so that the specific primer in the mixture is favored and a limited length of<br />

primer extension is achieved. In the second step, dideoxynucleotide terminations are<br />

effected at a more stringent elevated annealing/elongation temperature. The result is<br />

that only the extended and labeled primers enter into the termination reactions.<br />

We have used this method to sequence amplified cytochrome p450 cDNA fragments<br />

with a highly degenerate inosine-containing primer (1,2). In our case, a set of degenerate<br />

primers was used to amplify a presumably novel cytochrome p450 gene(s). The upstream<br />

sense primer was a mix of 192, 20mer, containing three inosines that theoretically could<br />

anneal to 12,288 different sequences. The downstream antisense primer was a mix of<br />

144, 23 mer, containing five inosines or 147,456 different possible sequences.<br />

In what follows, we will only describe the sequencing reactions. Procedures for<br />

sequencing gel electrophoresis can be found in Chapter 51.<br />

2. Materials<br />

1. A thermal cycler: Cetus Perkin–Elmer Model 480 (see Note 1).<br />

2. PCR tubes (0.5 mL).


370 Shen et al.<br />

3. Mineral oil.<br />

4. Gel-purification kits/reagents, such as QIAEX Gel Extraction Kit (Qiagen #20020,<br />

Chatsworth, CA) or QiaOuick Gel Extraction Kit (Qiagen #28704).<br />

5. All buffers and solutions must be free of DNase.<br />

6. α- 35 S-dATP (10 µCi/µL 1000 Ci/mmol; Amersham Corp).<br />

7. Sequencing primers (degenerate primers) dissolved in H 2 O, or 0.1× TE buffer.<br />

The sequencing reaction reagents can be homemade. However, we recommend purchasing<br />

them from a commercial company to ensure uniform performance. In the following<br />

materials, we include the catalog number for US <strong>Bio</strong>chemicals (Cleveland, OH).<br />

8. Reaction buffer (USB #71030): 260 mM Tris-HCl, pH 9.5; 65 mM MgCl 2 .<br />

9. ∆Taq DNA polymerase (USB #71059) or Taq DNA polymerase, (USB# 71057): 32 U/µL.<br />

10. Taq DNA polymerase dilution buffer (USB #71051): 10 mM Tris-HCl, pH 8.0, 1 mM<br />

2-mercaptoethanol, 0.5% Tween-20, and 0.5% Nonidet P-40.<br />

11. Four separate primer label mixes: dGTP label mix: 3.0 µM (USB #71034); dATP label<br />

mix: 3.0 µM (USB #71036); dTTP label mix: 3.0 µM (USB #71037); and dCTP label<br />

mix: 3.0 µM (USB #71038).<br />

12. Four separate termination mixes: ddG terminator mix: 15 µM each dGTP, dATP, dTTP,<br />

dCTP, and 22.5 µM ddGTP (USB #71020); ddA termination mix: 15 µM each dGTP, dATP,<br />

dTTP, dCTP, and 300 µM ddATP (USB #71035); ddT termination mix: 15 µM each dGTP,<br />

dATP, dTTP, dCTP, and 450 µM ddTTP (USB #71040); and ddC terminator mix: 15 µM<br />

each dGTP, dATP, dTTP, dCTP, and 75 µM ddCTP (USB #71025).<br />

13. Stop/gel-loading solution (USB #70724): 95% formamide, 20 mM EDTA, 0.05% bromophenol<br />

blue, and 0.05% xylene cyanol FF.<br />

14. 1× TE buffer: 10 mM Tris-HCl, pH 8.0, 1 mM EDTA.<br />

3. Methods<br />

3.1. Preparation of DNA as a Sequencing Template (see Note 2)<br />

1. After gel electrophoresis, PCR fragment(s) of interest is cut out from the gel.<br />

2. DNA in the gel is purified with the Qiagen gel-purification kit, and final PCR products are<br />

resuspended in a proper amount of 0.1× TE buffer.<br />

3. To estimate the amount of PCR product, run an aliquot of the PCR products on an agarose<br />

gel. The amount of DNA may be estimated by a comparison with the amount of DNA<br />

that was used in the mol-wt ladder.<br />

In the following steps, always keep tubes on ice, unless otherwise indicated.<br />

4. Prepare the following labeling PCR mix (see Note 3):<br />

H 2 O 10–8 µL<br />

DNA (in 0.1× TE) (need total of 25–100 ng) 11–9 µL<br />

Reaction buffer 12 µL<br />

Degenerate primers (5–200 µM) 11 µL<br />

dGTP label mix 11 µL<br />

dCTP label mix 11 µL<br />

dTTP label mix 11 µL<br />

α- 35 S-dATP (10 µCi/µL >1000 Ci/mmol) 10.5 µL<br />

Taq DNA polymerase (4 U/µL) (diluted in Taq dilution buffer) 12 µL<br />

Total volume 17.5 µL<br />

Cover the label PCR mix with 10–20 µL of mineral oil.<br />

3.2. Labeling PCR (see Note 4)<br />

1. Run the following PCR program: presoak at 94°C for 3 to 5 min followed by 45 cycles<br />

of 95°C for 30 s and 52°C for 30 s.


Inosine-Containing Primers 371<br />

2. Transfer 15 to 16 µL of the above labeled mixture to a new tube. Avoid carryover of any<br />

mineral oil. This can be done easily by putting the pipetting tip directly below the oil<br />

without touching the wall of the tube.<br />

3. Optional (see Note 5): Load 1 to 2 µL with 1 µL of gel-loading buffer to a sequencing<br />

gel to check the labeling efficiency.<br />

4. Termination PCR mix: For each of the labeled mixes, prepare four tubes labeled as “G,”<br />

“A,” “T,” and “C.” To each of the tubes, add 4 µL of termination mix G, A, T, or C (this can<br />

be done toward the end of label-PCR procedure). Add 3.5 µL of the label mix to each of<br />

the tubes. Cover the termination PCR mix with 8 to 10 µL of mineral oil.<br />

5. Termination PCR: Cycle between 95°C for 30 s and 72°C for 90 s (see Note 6).<br />

6. While the termination PCR is under way, prepare four clean 0.5-mL tubes labeled G, A, T,<br />

or C. To each of them add 4 µL of stop/gel-loading solution.<br />

7. Transfer 6 to 7 µL of termination mix to these tubes with the stop/loading solution.<br />

Avoid carryover of mineral oil. Mix and spin down briefly and store at –20°C (good for<br />

up to 1 mo). These samples are ready for the sequencing gel (use 3 µL to load a gel).<br />

See Chapter 51.<br />

8. Sequencing results: run sequencing gel, perform autoradiography, and read the sequence<br />

(see Note 7).<br />

4. Notes<br />

1. Cetus Perkin–Elmer thermal cycler Model 9600 also may be used. If it is, use 0.1-mL<br />

tubes; no mineral oil on the top of the reaction solution is needed. The PCR program<br />

should be adjusted accordingly in the procedure.<br />

2. Other methods of DNA preparation are also acceptable as long as “clean” DNA is<br />

obtained.<br />

3. Other {a- 35 S-labeled nucleotides may also be used, but the label mix must be changed<br />

accordingly. The concentration of the stock of degenerate primers in the reaction is<br />

dependent on the degree of degeneracy. In our case, the stock concentration of our >100∞<br />

degenerate primer was 200 µM. Because radioactive 35 S is used for these experiments,<br />

always be careful and follow the safety operation procedure for your institute. Check<br />

with your radiation safety officer for the authorized amount of radioactivity that you can<br />

handle at any one time.<br />

4. Depending on the sequencing primer, the annealing/elongation temperature or time may<br />

have to be optimized to give proper primer extension and labeling. The purposes of the<br />

label PCR is to have a sufficient amount of specific primer in the primer mix to anneal to a<br />

specific site on the DNA template and to extend the annealed primer for a limited nucleotide<br />

length with Taq DNA polymerase. The first purpose can be achieved by choice of an optimal<br />

annealing temperature and/or time. In our case (200 pmol of the 20 mer with inosine and<br />

a degeneracy of more than 100∞), we used 52°C and 30 s. Depending on circumstances,<br />

this temperature and the annealing time may need to be adjusted. The method of generating<br />

a limited elongated primer plus labeling of the primer is accomplished by using a shorter<br />

annealing/elongation time at a suboptimal temperature (for Taq activity), but still a stringent<br />

temperature for annealing and a low dNTP concentration. In this way, it is not necessary to<br />

know the sequence of the downstream flanking region of the sequencing primer.<br />

5. Loading 1 to 2 µL of labeled mix to run a gel to check that the length of primer extension<br />

and label efficiency is optional. This can be run along with the sequencing sample after<br />

all the reactions are finished. Using our p450 degenerate sequencing primers under these<br />

labeling PCR conditions, we obtained an average primer extension of 15 to 25 bp.<br />

6. The temperature used for both the annealing of labeled/extended primers to the DNA<br />

template for the elongation/termination reaction was 72°C. Only the prelabeled and pre-


372 Shen et al.<br />

extended primers, which are the specific primers in the primer mix, would be allowed<br />

during the termination/elongation because of the elevated temperature. If more template<br />

DNA is available, fewer cycles may be used.<br />

7. Depending on the length of labeled primers, the readable sequence will vary. For our<br />

case of highly degenerate inosine-containing primers of p450 genes (see Subheading 1.<br />

for a description of our p450 primers) a ladder from 25 bp downstream of the primer<br />

was readable up to 300 bp.<br />

8. A similar protocol of this method would be to omit one of the four dNTPs in the label step<br />

and use at least one α- 35 S-labeled dNTP in the labeling mix. This will give an incomplete<br />

elongation of the sequencing primer during the labeling step because the primer extension<br />

will stop at the proper position when the omitted nucleotide is not present. The elongated<br />

primers may be labeled if the labeled nucleotide is by chance present between the sequence<br />

primer and the omitted nucleotide. This method is useful to sequence DNA when some<br />

sequence <strong>info</strong>rmation immediately downstream from the sequencing primer is available. In<br />

such a case, one can decide which nucleotide to omit or to label in the label mix.<br />

References<br />

1. Shen, Z., Liu, J., Wells, R. L., and Elkind, M. M. (1993) Cycle sequencing using degenerate<br />

primers containing inosines. <strong>Bio</strong>Techniques 15, 82–89.<br />

2. Shen, Z., Wells, R. L., Liu, J., and Elkind, M. M. (1993) Identification of a cytochrome<br />

p450 gene by reverse transcription-PCR using degenerate primes containing inosines. Proc.<br />

Natl. Acad. Sci. USA 90, 11,483–11,487.<br />

3. Shen, Z., Liu, J., Wells, R. L., and Elkind, M. M. (1994) cDNA cloning, sequence analysis,<br />

and induction by aryl hydrocarbons of a murine cytochrome p450 gene, Cyplbl. DNA<br />

Cell <strong>Bio</strong>l. 13, 763–769.<br />

4. Shen, Z., Denison, K., Lobb, R., Gatewood, J., and Chen, D. J. (1995) The human and<br />

mouse homologs of yeast RAD52 genes: cDNA cloning, sequence analysis, assignment<br />

to human chromosome 12pl2.2-pl3, and mRNA expression in mouse tissues. Genomics<br />

25, 199–206.<br />

5. Compton, T. (1990) Degenerate primers for DNA amplification, in PCR Protocol, a Guide<br />

to Methods and Applications (Innis, M. A., Gelfand, D. H., Sninsky, J. J., and White, T. J.,<br />

eds.), Academic, San Diego, CA, pp. 39–45.<br />

6. Lee, C. C. and Caskey, C. T. (1990) cDNA cloning using degenerate primers, in PCR<br />

Protocol, a Guide to Methods and Applications (Innis, M. A., Gelfand, D. H., Sninsky, J. J.,<br />

and White, T. J., eds.), Academic, San Diego, CA, pp. 46–59.<br />

7. Knoth, K. S., Roberds, S., Poteet, C., and Tamkun, M. (1988) Highly degenerate inosinecontaining<br />

primers specifically amplify rare cDNA using the polymerase chain reaction.<br />

Nucleic Acids Res. 16, 10932.<br />

8. Erlich, H. A., Gelfand, D., and Sninsky, J. J. (l99l) Recent advances in the polymerase<br />

chain reaction. Sciences 252, 1643–1651.<br />

9. Murray, V. (1989) Improved double strand DNA sequencing using the linear polymerase<br />

chain reaction. Nucleic Acids Res. 17, 8889.<br />

10. Smith, D. P., Jonstone, E. M., Little, S. P., and Hsiung, H. M. (1990) Direct DNA sequencing<br />

of cDNA inserts from plaques using the linear polymerase chain reaction. <strong>Bio</strong>techniques<br />

9, 48–52.


Unknown Genomic Sequences 373<br />

55<br />

Determination of Unknown Genomic Sequences<br />

Without Cloning<br />

Jean-Pierre Quivy and Peter B. Becker<br />

1. Introduction<br />

The inherent problems of sensitivity and specificity that one encounters when trying<br />

to determine a particular nucleotide sequence directly in its genomic context can be<br />

overcome by selective amplification of the region of interest. This amplification of<br />

the target DNA is usually achieved by one of two strategies: The relevant piece of<br />

DNA may be cloned and therefore amplified in a bacterial cell or, alternatively, the<br />

desired fragment may be amplified in vitro using PCR technology. Both strategies have<br />

drawbacks. The cloning of a specific genomic sequence is labor intensive, lengthy,<br />

and sometimes even difficult to achieve. The PCR amplification requires that enough<br />

sequence <strong>info</strong>rmation is known to be able to design the two specific amplification<br />

primers and is therefore limited to sequencing alleles of already-known DNA. There<br />

are, however, many cases that would benefit from the determination of unknown<br />

genomic sequence close to a known piece of DNA. With a particular cDNA in hand, one<br />

may wish, for example, to determine genomic gene sequences, such as the promoter<br />

of the gene, its introns, or 5′- and 3′-nontranscribed regions. The protocol presented<br />

here uses ligation-mediated polymerase chain reaction (LM-PCR) to amplify unknown<br />

genomic DNA next to a short stretch (about 100 bp) of known sequence and details<br />

a convenient procedure to determine the new sequence by dideoxy sequencing (1).<br />

The procedure may form the basis for “walking sequencing” strategies to determine<br />

large regions of continuous sequence <strong>info</strong>rmation starting from a limited piece of<br />

known DNA.<br />

The central feature of the LM-PCR technique is the ligation of a known short<br />

oligonucleotide, the “linker,” to selected ends of genomic DNA fragments (Fig. 1).<br />

These generic linker sequences provide the second primer for amplification of linked<br />

fragments in combination with an oligonucleotide based on known sequences. LM-PCR<br />

was first introduced for genomic footprinting and chemical sequencing (2,3). The<br />

disadvantages of chemical sequencing over the chain termination method (1) prompted<br />

us to adapt LM-PCR technology for direct dideoxy sequencing of genomic DNA (4).<br />

The underlying procedure is derived from a variation of the original LM-PCR protocol<br />

called “Linker Tag Selection LM-PCR” (5).<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

373


374 Quivy and Becker<br />

Fig. 1. The use of Linker Tag Selection LM-PCR for genomic dideoxy sequencing. Bold<br />

lines denote known sequences, and thin lines the unknown sequences to be determined. The<br />

arrows and dotted lines indicate primer extension reactions. R: Strategic restriction site. P1, P2,<br />

P3, and LP stand for the primers 1, 2, 3, and the linker primer, respectively. The biotin moiety on<br />

the linker is represented by a filled circle, and the streptavidin-coated paramagnetic beads by the<br />

shaded boxes. (H): Radiolabeled P3 is used to prime dideoxy sequencing reactions.<br />

The steps of the reaction are outlined in Fig. 1. A restriction enzyme is selected<br />

that cleaves the genomic DNA somewhere within the unknown sequence but within<br />

1 kb from the known sequence for which the specific primers have been designed.<br />

Cleavage creates a defined end (A). The cleaved genomic DNA is denatured, and<br />

the gene-specific primer 1 is annealed to the known sequence and extended by a<br />

polymerase until the end of the restriction fragment (B). This creates a blunt end to<br />

which a short double-stranded linker is ligated (C). The linker DNA consists of two<br />

complementary oligonucleotides, the longer one being biotinylated at its 5′-end. After<br />

denaturation, the specific primer 2, representing sequences more 3′ from primer 1 on<br />

the lower strand of the known sequence, is then annealed to the upper strand, which<br />

now carries the biotinylated linker oligonucleotide at its 5′-end. The primer is again<br />

extended to the end (D) creating a double-stranded fragment that contains a region<br />

of unknown DNA flanked by known sequences. The genomic sequences can now be


Unknown Genomic Sequences 375<br />

Fig. 2. Genomic sequences of the Drosophila hsp27 promoter region obtained following the<br />

outlined procedure. Arrows indicate the full-length restriction fragment. (A) Long and short<br />

gel runs of the same sequencing reaction obtained using the enzyme NruI, which cut 821 bases<br />

upstream of the 5′-end of primer 1. (B) Upper part of a sequence obtained using the enzyme PstI<br />

cutting 332 bases away from the 5′-end of primer 1. The sequence of the linker oligonucleotide<br />

at the end of the genomic sequence can be unambiguously identified.<br />

amplified by PCR using a combination of the biotinylated linker primer and primer<br />

2 (E). The biotinylated amplification products are then immobilized on streptavidincoated<br />

paramagnetic beads and purified from the PCR in a magnetic field (F). This<br />

step efficiently removes unincorporated primers, which interfere with the subsequent<br />

sequencing reaction. The immobilization also facilitates the handling of the template<br />

fragments during the subsequent steps and specifically allows the sequencing of a singlestranded<br />

template. The immobilized fragments are denatured, and the complementary<br />

strand is removed by washes (G). Radioactively labeled specific primer 3, again located<br />

3′ to primer 2 on the lower strand, now serves to prime a standard chain termination<br />

sequencing reaction that finally generates the sequencing ladder (H). Sequences start to<br />

be readable about 25 bases 3′ from primer 3, and may extend for up to 1 kb depending on<br />

the location of the restriction site and the efficiency of the overall process (Fig. 2A).<br />

Important features of the procedure are the use of a proofreading polymerase for the<br />

PCR amplification (the Vent DNA polymerase possesses a 3′-5′ exonuclease activity)


376 Quivy and Becker<br />

that decreases the error rate during the amplification, and the use of an enzyme without<br />

the 3′-5′ exonuclease activity (e.g., Vent Exo) for efficient single primer extensions.<br />

The immobilization of the amplified fragments on paramagnetic beads exploiting the<br />

strong streptavidin/biotin interaction is crucial for the efficiency of the sequencing<br />

reaction because it allows the efficient removal of interfering primers from the PCR and<br />

permits the use of a single-strand template (4,5). The attachment of the sequenced DNA<br />

fragment to the solid support does not create a steric hindrance for the polymerase.<br />

Frequently, the sequence of the linker oligonucleotide itself can be determined at the<br />

very end of the genomic sequence (see Fig. 2B).<br />

The length of sequence determined critically depends on the ability to resolve long<br />

fragments with single nucleotide resolution, provided that the restriction site used for<br />

linker ligation is not too close to the known sequence. Fragments of over 800 bp have<br />

yielded reliable sequence <strong>info</strong>rmation (Fig. 2A). Longer sequences can be obtained<br />

in walking strategies where the newly determined sequence is in turn used to design<br />

further reaching sets of primers. The presented strategy critically relies on the previous<br />

identification of a suitable restriction site, ideally between 0.5 and 1 kb away from<br />

the known sequence. Too short fragments will yield little new sequence <strong>info</strong>rmation,<br />

whereas large fragments (exceeding 1 kb) do not work, presumably because of<br />

decreasing efficiencies in DNA denaturation and primer extensions reactions. Because<br />

any kind of restriction enzyme will work, a site can be conveniently identified on a<br />

Southern blot testing a small selection of enzymes that cut the genome at a reasonable<br />

frequency. Alternatively, a selection of enzymes can simply be tried at random in an<br />

LM-PCR sequencing reaction. To increase the chances that the reaction will work, the<br />

genomic DNA may be cleaved with a whole cocktail of enzymes that collectively have<br />

a high likelihood to produce a suitable restriction fragment.<br />

2. Materials<br />

2.1. Purification and Restriction of Genomic DNA<br />

1. Suspension of nuclei from desired organism (see Note 1).<br />

2. 0.5 M EDTA, pH 8.0.<br />

3. RNase A, DNase free, 10 mg/mL (Boehringer, Mannheim).<br />

4. Aqueous solution of N-lauroylsarkosine (sarkosyl), 20% (P/V) (Sigma).<br />

5. Proteinase K, 10 mg/mL (Merck).<br />

6. Phenol, highest quality, neutralized, and equilibrated with TE (10 mM Tris-HCl, pH 7.5;<br />

1 mM EDTA; Aurresco).<br />

7. Phenol/chloroform/isoamyl alcohol mixture (25241; Aurresco).<br />

8. Chloroformisoamyl alcohol (241; Merck).<br />

9. 0.3 M sodium acetate, pH 5.2.<br />

10. Ethanol 100%.<br />

11. Ethanol 80%.<br />

12. TE: 10 mM Tris-HCl, pH 7.5, 1 mM EDTA.<br />

13. Restriction enzyme with suitable 10× reaction buffer (see Note 2).<br />

2.2. Primer Design, Primer Purification, and Annealing<br />

of the Linker Primer (see Note 3)<br />

The primers were synthesized on an ABI 394 DNA synthesizer and gel purified<br />

(see Subheading 3.2.).


Unknown Genomic Sequences 377<br />

1. Primer 1 (P1): a 18-22 mer with a calculated T m around 45 to 50°C. Working concentration:<br />

0.5 pmol/µL.<br />

2. Primer 2 (P2) should have the same melting temperature as the long-linker primer (see<br />

Note 4) used for the PCR amplification, in our case a 25 to 27 mer with a calculated T m<br />

of 60 to 65°C. The guanine-cytosine content is usually between 45 and 55%. It does not<br />

need to overlap with P1, but a 5-bp overlap has worked. It should be internal to primer 2 to<br />

increase the specificity of the overall reaction. Working concentration: 10 pmol/µL.<br />

3. Primer 3 (P3): an 18 to 22 mer with a calculated T m around 45 to 50°C, but longer<br />

oligos with higher GC contents will also work. It should be internal to P2 to increase the<br />

specificity. It will be 32 P kinased with an SA of 10 7 cpm/pmol (see Subheading 3.3.).<br />

Working concentration: 0.2 pmol/µL.<br />

4. Long-linker oligonucleotide: 5′ CACCCGGGAGATCTGAATTC 3′ (see Note 4). It is<br />

biotinylated at its 5′-end during synthesis by incorporation of a biotin-2-o-propylphosphoramidite.<br />

It should be unphosphorylated.<br />

5. Short-linker oligonucleotide: 5′ GAATTCAGATC 3′, dephosphorylated.<br />

6. Oligo loading mix: 10% glycerol in formamide.<br />

7. Denaturing polyacrylamide gel: 14.5% acrylamide, 0.5% bis-acrylamide, 7 M urea, 1× TBE.<br />

Size: 25 × 25 × 0.1 cm.<br />

8. Formamide loading buffer: 96% formamide, 0.05% xylene cyanol, 0.05% bromophenol<br />

blue, 10 mM EDTA.<br />

9. PE buffer: 50 mM Tris-HCl, pH 7.5, 100 mM NaCl, 1 mM EDTA, 1% phenol (v/v).<br />

10. Chloroformisoamyl alcohol (241; Merck).<br />

11. 5 M LiCl.<br />

12. 1 M MgCl 2 .<br />

13. Ethanol 100%.<br />

14. Ethanol 80%.<br />

15. TBE: 90 mM Tris-borate, 1 mM EDTA, pH 8.3.<br />

2.3. Kinasing of Primer 3<br />

1. P3 at 10 pmol/µL.<br />

2. Polynucleotide kinase buffer 10×: 700 mM Tris-HCl, pH 7.6, 100 mM MgCl 2 , 50 mM DTT.<br />

3. γ 32 PATP, 5000 Ci/mmol (Redivue, Amersham).<br />

4. Polynucleotide kinase 10 U/µL (NE <strong>Bio</strong>labs).<br />

5. 50 mM EDTA, pH 8.0.<br />

6. G-25 fine spin columns (Boehringer, Mannheim).<br />

2.4. First Primer Extension<br />

1. P1 at 0.5 pmol/µL in TE, pH 7.5.<br />

2. 1 N NaOH.<br />

3. TES buffer: 560 mM TES, free acid (Sigma), 240 mM HCl, 100 mM MgCl 2 .<br />

4. Vent buffer (10×): 100 mM KCl; 100 mM (NH 4 ) 2 SO 4 , 200 mM Tris-HCl, pH 8.8, 20 mM<br />

MgSO 4 , 0.1% Triton X-100 (NE <strong>Bio</strong>labs).<br />

5. dNTP solution: 10 mM dNTPs, (Boehringer, Mannheim); keep in small aliquots at –20°C.<br />

Do not freeze/thaw more than three times.<br />

6. Vent Exo DNA polymerase, 2 U/µL (NE <strong>Bio</strong>labs).<br />

2.5. Ligation<br />

1. Ligase buffer (10×): 500 mM Tris-HCl, pH 7.5, 100 mM MgCl 2 , 100 mM DTT, 10 mM<br />

ATP, 250 µg/mL BSA (NE <strong>Bio</strong>labs).


378 Quivy and Becker<br />

2. Solution of 40% PEG 8000 (Sigma), filtered through a 0.22-µm filter. It will take some<br />

time to dissolve the PEG in water. Incubate at room temperature on a rotating wheel for<br />

several hours. It also takes some force to filter the solution using a syringe.<br />

3. T4 DNA ligase, 400 U/µL (NE <strong>Bio</strong>labs).<br />

4. Annealed linker (see Subheading 3.4.).<br />

2.6. PCR<br />

1. TE: 10 mM Tris-HCl, pH 8.5; 1 mM EDTA, pH 8.5.<br />

2. Phenolchloroformisoamyl alcohol (25241; Aurresco).<br />

3. Solution of 7.5 M ammonium acetate containing 25 µg/mL yeast tRNA (Boehringer<br />

Mannheim). Crude yeast tRNA has to be cleaned by multiple organic extractions and<br />

ethanol precipitation.<br />

4. Vent buffer (see Subheading 2.4.).<br />

5. dNTP solution (see Subheading 2.4.).<br />

6. 100 mM MgSO 4 .<br />

7. P2 solution, 10 pmol/µL (see Subheading 2.2.).<br />

8. Long-linker primer, 10 pmol/µL (see Subheading 2.2.).<br />

9. Vent DNA polymerase 2 U/µL (NE <strong>Bio</strong>labs).<br />

10. Perkin–Elmer thermal cycler.<br />

11. PCR tubes (Perkin–Elmer).<br />

12. Mineral oil (PCR-grade).<br />

2.7. Tag Selection of the PCR Products<br />

1. Dynabeads M-280 streptavidin (Dynal, Oslo, 10 mg/mL).<br />

2. Magnetic particle concentrator (MPC, Dynal).<br />

3. Phosphate-buffered saline (PBS), pH 7.4.<br />

4. PBS, pH 7.4; 0.01% BSA (molecular biology grade).<br />

5. BW solution: a 11 mixture of TE and 5 M NaCl.<br />

2.8. Sequencing Reaction (see Note 7)<br />

1. MPC.<br />

2. 150 mM NaOH, freshly prepared.<br />

3. TE, pH 7.5.<br />

4. Vent buffer (see Subheading 2.4.).<br />

5. Termination mixes made up in 1× vent buffer:<br />

A-mix: 900 µM ddATP, 30 µM dATP, 100 µM dCTP, 100 µM dGTP, 100 µM dTTP.<br />

C-mix: 480 µM ddCTP, 30 µM dATP, 37 µM dCTP, 100 µM dGTP, 100 µM dTTP.<br />

G-mix: 400 µM ddGTP, 30 µM dATP, 100 µM dCTP, 37 µM dGTP, 100 µM dTTP.<br />

T-mix: 720 µM ddTTP, 30 µM dATP, 100 µM dCTP, 100 µM dGTP, 33 µM dTTP.<br />

6. Labeled P3 (see Subheading 3.3.).<br />

7. Circumvent sequencing buffer 10×: 100 mM KCl, 100 mM (NH 4 ) 2 SO 4 , 200 mM Tris-HCl,<br />

pH 8.8, 50 mM MgSO 4 (NE <strong>Bio</strong>labs).<br />

8. Triton X-100 solution, 3% (v/v) in water.<br />

9. Vent Exo DNA polymerase, 2 U/µL (NE <strong>Bio</strong>labs).<br />

10. Formamide loading buffer (see Subheading 2.2.).<br />

11. Sequencing gel.<br />

12. Fixing solution: 10% acetic acid, 10% methanol.<br />

13. Dupont NEN reflection films and corresponding cassettes.


Unknown Genomic Sequences 379<br />

3. Methods<br />

3.1. Purification of Genomic DNA and Restriction (see Note 1)<br />

1. Isolate nuclei from cells of interest by suitable methods (see Note 1) and spin them down<br />

to obtain the nuclear pellet.<br />

2. Resuspend the pellet in 1 mL of 0.5 M EDTA. Avoid harsh vortexing and vigorous pipetting<br />

to prevent shearing (see Note 1).<br />

3. Add 25 µL of RNase A and 25 µL of Sarkosyl, mix by inverting the tube, and incubate<br />

for 3 h at 37°C on a rotating wheel.<br />

4. Add 25 µL of proteinase K, mix by inverting the tube, and incubate overnight at 37°C<br />

on a rotating wheel.<br />

5. Add 1 mL of phenol and mix by inverting the tube several times. Spin to separate the<br />

phases and collect the lower phase and interphase (the lower phase is the aqueous phase<br />

owing to the high density of 0.5 M EDTA).<br />

6. Repeat step 5, but do not take the interphase.<br />

7. Add 1 mL of phenolchloroform and collect the upper phase (which is now the aqueous<br />

phase).<br />

8. Dialyze overnight (or longer) against TE, pH 7.5, with at least four changes of TE. Avoid<br />

a large increase in volume by keeping the dialysis bag tight.<br />

9. Precipitate DNA with one-tenth vol of 0.3 M Na acetate and 2.5 vol of cold absolute<br />

ethanol.<br />

10. Spin at 4°C to collect DNA pellet, wash with 80% ethanol, remove residual ethanol, but<br />

do not dry too long since the DNA will be difficult to redissolve.<br />

11. Dissolve DNA in TE, pH 7.5, and store at 4°C.<br />

12. Restriction digest of genomic DNA: Digest 10 µg of DNA with 1 U/µg of restriction<br />

enzyme according to the manufacturer’s recommendation (see Notes 2 and 4) for at least<br />

3 h (up to overnight).<br />

13. Extract DNA once with phenol, once with phenol:chloroform, once with chloroform, and<br />

precipitate with one-tenth vol of 0.3 M Na acetate and 2.5 vol of cold 100% ethanol.<br />

14. Spin at 4°C, wash DNA pellet with 80% ethanol, and remove residual ethanol in the<br />

SpeedVac. Again, do not dry too long. Resuspend DNA in TE, pH 7.5, and adjust<br />

concentration to 1 µg/µL (OD at Aµ260).<br />

15. Extract once more with chloroform, and remove traces of chloroform in the SpeedVac.<br />

3.2. Purification of Oligonucleotide Primers and Annealing<br />

of the Linker Fragment<br />

1. Dry down 75 nmol of the oligonucleotide in the SpeedVac concentrator and dissolve in<br />

75 µL of oligo loading mix. Heat for 5 min at 75°C and load 5∞15 µL onto a prerun<br />

denaturing polyacrylamide gel. In a separate slot load some formamide-loading buffer<br />

to monitor the run. Electrophorese until the bromophenol marker dye migrates to two<br />

thirds of the gel.<br />

2. By ultraviolet shadowing (6), locate the band corresponding to the full-length oligonucleotide<br />

and excise from the gel using a razor blade.<br />

3. Transfer the polyacrylamide gel slice to a 1.5-mL reaction tube containing 1 mL of PE<br />

buffer and incubate overnight at 37°C.<br />

4. Filter the supernatant through a 0.22-µm filter, prewetted with PE, with the help of a<br />

2-mL syringe.<br />

5. Wash another 100 µL of PE buffer through the filter.<br />

6. Extract the pooled PE solutions with chloroform and dispense 2∞450 µL of the upper<br />

phase into fresh tubes.


380 Quivy and Becker<br />

7. Precipitate the oligonucleotides by the addition of 36 µL of 5 M LiCl, 4.5 µL of 1 M<br />

MgCl 2 , and 1 mL cold 100% ethanol.<br />

8. Mix by vortexing and let precipitate at –70°C for 15 min.<br />

9. Spin at 4°C for 20 min in a tabletop centrifuge, wash the pellet with 80% ethanol, dry in<br />

the SpeedVac, and resuspend the oligonucleotide in TE, pH 7.5.<br />

10. Determine the concentration (OD at A 260 ) and dilute some aliquots to the working<br />

concentration.<br />

11. Annealing of linker oligos: combine in a 1.5-mL reaction tube: 20 pmol/µL of each linker<br />

oligonucleotide in 250 mM Tris (pH 7.5), 5 mM MgCl 2 . Heat at 95°C for 5 min, transfer<br />

to a beaker containing boiling water, and allow to cool slowly at room temperature for 5 h<br />

(up to overnight in the cold room). Aliquot the linker solution and store at –20°C. Aliquots<br />

are only used once and never refrozen.<br />

3.3. Kinasing of Primer 3<br />

1. Combine in a tube 1 µL of P3 (10 pmol), 3 µL of 10× T4 polynucleotide kinase, and<br />

10.5 µL of water.<br />

2. Add 15 µL of γ 32 PATP and 0.5 µL of polynucleotide kinase.<br />

3. Incubate for 30 min at 37°C.<br />

4. Add 20 µL of 50 mM EDTA, pH 8.0, and heat at 65°C for 10 min.<br />

5. Purify the labeled oligonucleotide from the unincorporated label by a G-25 spin column.<br />

3.4. First Primer Extension<br />

1. Combine in a tube: 0.5 to 1 µg of restricted genomic DNA (see Subheading 3.1. and Note 4),<br />

1 µL P1 (0.5 pmol), 1 µL 1 N NaOH, and water to 8 µL.<br />

2. Incubate at 65°C for 5 min.<br />

3. Immediately add 2 µL of TES buffer and mix.<br />

4. Spin to collect and incubate for 10 min at room temperature.<br />

5. Add 9 µL of a mix containing 2 µL of 10× Vent buffer, 0.4 µL of 10 mM dNTPs,<br />

6.6 µL of H 2 O.<br />

6. Incubate at 50°C for 10 to 20 min.<br />

7. Add 1 µL of the Vent Exo (2 U).<br />

8. Incubate for 10 min at 76°C.<br />

9. Chill on ice, spin to collect liquid, and proceed immediately to ligation.<br />

3.5. Ligation<br />

1. Prepare a premix containing 5 µL of 10× ligase buffer, 19 µL of 40% PEG 8000, and<br />

5 µL of annealed linker (20 pmol/µL). Mix well by pipetting since the resulting solution<br />

is very viscous.<br />

2. Add 29 µL of this premix to the first primer extension reaction (see Section 3.4.).<br />

3. Add 1 µL of T4 DNA ligase (400 U), mix well by pipetting, and incubate overnight<br />

at 17°C.<br />

3.6. PCR<br />

1. To the ligation reaction, add 150 µL of TE, pH 8.5, and mix by vortexing.<br />

2. Add 150 µL of phenolchloroformisoamyl alcohol (25241) mix by vortexing, and<br />

then spin for 5 min.<br />

3. Collect the upper aqueous phase and transfer to a tube containing 10 µL of 7.5 M<br />

NH 4 Ac/yeast tRNA. Add 750 µL of cold 100% ethanol, mix by vortexing, and let<br />

precipitate for 15 min at –20°C.


Unknown Genomic Sequences 381<br />

4. Centrifuge for 20 min at 4°C, remove the supernatant, wash the pellet with 500 µL of 80%<br />

ethanol, and dry in the SpeedVac.<br />

5. Dissolve DNA in 20 µL of water, transfer to a 500-µL PCR tube, and keep on ice.<br />

6. Prepare a premix containing 5 µL of 10× Vent buffer, 1 µL of 10 mM dNTPs, 1 µL of<br />

100 mM MgSO 4 (see Note 5), 1 µL of P2 solution, 1 µL of long-linker primer, and 19.5 µL<br />

of water.<br />

7. Immediately before use, add 1.5 µL of Vent DNA polymerase (3 U) to the premix and<br />

mix by pipetting. Add premix to the PCR tube containing the DNA, then add two drops<br />

of mineral oil, and keep on ice.<br />

8. Transfer the tube to the thermal cycler (one droplet of mineral oil in sample holders)<br />

preheated to 95°C, incubate for 2.5 min at 95°C, and subject to 18 cycles as follows: 95°C<br />

for 1 min, 60°C for 2 min (see Note 5), and 76°C for 3 min. Allow an increase of 5 s/cycle<br />

for the extension step, and end the cycling by a 10-min incubation at 76°C.<br />

9. The PCR can be used immediately for the labeling step or stored frozen at –20°C.<br />

3.7. Linker Tag Selection of the PCR Product<br />

1. Resuspend the Dynabeads M-280 streptavidin well and take 50 µL (500 µg beads) into<br />

a 1.5-mL reaction tube.<br />

2. Concentrate in the magnetic rack (MPC, Dynal), and remove the supernatant (0.01%<br />

BSA; see Note 6).<br />

3. Wash the beads with 100 µL of PBS (pH 7.4), concentrate, and remove the supernatant.<br />

Repeat this step.<br />

4. Resuspend beads in 100 µL of PBS/0.01% BSA by pipetting up and down until a<br />

homogenous suspension is achieved. Concentrate and remove supernatant.<br />

5. Wash the beads with 100 µL of BW solution by pipetting up and down until no aggregates<br />

are seen. Concentrate again. Repeat this step.<br />

6. Finally resuspend the beads in 100 µL of BW solution. The beads are now ready to be<br />

used.<br />

7. Add the PCR to the beads and mix well. Avoid mineral oil, which spoils the magnetic<br />

separation of beads (we do not recommend a chloroform extraction because traces of this<br />

solvent adversely affects later steps in the reaction).<br />

8. Incubate the tube on a rotating wheel at room temperature for 30 min. Assure that the beads<br />

do not sediment in the tube but also avoid a spreading of the liquid over the entire tube<br />

wall. A suitable agitation can be obtained by adjusting the rotation angle.<br />

9. Concentrate beads and discard supernatant.<br />

10. Wash the beads with 100 µL of BW solution.<br />

3.8. Template Denaturation and Sequencing Reaction (see Note 7)<br />

1. Resuspend beads well in 100 µL 150 mM NaOH by pipetting up and down. Incubate<br />

at room temperature for 5 min with occasional gentle agitation, and then for 2 min<br />

at 50°C.<br />

2. Spin shortly to collect condensed liquid and beads trapped in the lid. Concentrate beads.<br />

3. Remove the supernatant and resuspend the beads in 100 µL of 150 mM NaOH. Spin<br />

shortly to collect all the NaOH solution and concentrate beads.<br />

4. Discard supernatant, resuspend the beads in 100 µL of TE, pH 8.5, and concentrate again.<br />

Repeat this wash with TE.<br />

5. Resuspend the beads in 50 µL of Vent buffer and keep on ice.<br />

6. Prepare four tubes labeled A, C, G, and T containing 3 µL of the respective termination<br />

mixes (see Subheading 2.8.).


382 Quivy and Becker<br />

7. Prepare a premix containing 0.2 pmol labeled P3, 1.5 µL of 10× circumvent sequencing<br />

buffer, 1 µL of Triton X-100, and add water to 15 µL.<br />

8. Concentrate beads and then remove supernatant.<br />

9. Resuspend the beads in 15 µL of premix.<br />

10. Transfer 3.5 µL of this bead suspension into each tube containing a termination mix and<br />

mix by pipetting.<br />

11. Heat at 95°C for 10 s and then transfer to 50°C. Incubate for 10 to 20 min with occasional<br />

gentle resuspension.<br />

12. Add 1 µL of Vent Exo (2 U) to each tube, mix well, and immediately transfer to 76°C.<br />

13. Incubate for 10 min and then chill on ice.<br />

14. Spin to collect liquid, concentrate beads, and discard supernatant.<br />

15. Resuspend beads in 50 µL of water, spin shortly to collect liquid, and concentrate beads.<br />

16. Carefully remove all liquid and resuspend the beads in 4 µL of formamide loading<br />

buffer0.15 NaOH (21). Mix well to dissolve all aggregates (see Note 8).<br />

17. Leave for 5 min at room temperature.<br />

18. Incubate 3 min at 76°C, spin to collect liquid, and chill on ice.<br />

19. Concentrate beads and then transfer supernatant into a fresh tube on ice.<br />

20. Check with a hand monitor that most of the radioactivity is in the supernatant. The bead<br />

pellet always contains radioactivity owing to trapped P3. Usually, we do not re-extract<br />

the beads.<br />

21. Load on a prerun sequencing gel (see Chapter 51); reactions can be stored at –20°C.<br />

22. Fix gel in fixing solution, dry onto blotting paper, and expose the dried gel to X-ray film<br />

in the presence of an intensifying screen. Readable sequences are usually obtained after<br />

overnight exposure.<br />

4. Notes<br />

1. The genomic DNA used for genomic sequencing needs to be clean and undegraded. Any<br />

shearing of the DNA during preparation and handling before the first primer extension must<br />

be avoided. Nicks between the restriction site and the P1 priming site will be converted to<br />

blunt ends during the first primer extension and will give rise to a background of dominant<br />

bands in all four sequencing lanes. We detail here one particular protocol that consistently<br />

yields high molecular weight genomic DNA of good quality. In principle, other methods<br />

can be followed. In any case, we advise to start a DNA prep with a nuclei isolation. Some<br />

suggestions for how to prepare nuclei have been described (7,8).<br />

2. Enzymes that produce blunt ends, with 5′- or 3′-overhangs can be used because the<br />

fragment is anyway converted to a blunt end after the first primer extension.<br />

3. All oligonucleotides should be gel-purified (see Subheading 3.2.) and stored in water or<br />

TE, pH 7.5, at two concentrations: a stock solution (determined after the purification) and<br />

a diluted working solution that should not be frozen and thawed more than five times.<br />

The sequence-specific primers 1–3 need to be designed for each new sequencing project.<br />

The long and short linker oligonucleotides are constant. The distances between primers<br />

1/2 and 2/3 should not be longer than 10 bases; overlaps of up to 5 bp are tolerated,<br />

but not required.<br />

4. This protocol was used to sequence in the context of the Drosophila genome, which<br />

required changes from the original protocol described for mammalian DNA. If sequences<br />

are to be determined in the context of the 10 times more complex mammalian genome,<br />

a few parameters have to be adjusted: The amount of starting DNA used should be<br />

increased about five times, and the longer-linker oligos should be taken from the original<br />

procedure (3).


Unknown Genomic Sequences 383<br />

5. The annealing temperature and the optimal magnesium concentration may be optimized<br />

specifically for each primer, but the given conditions worked reasonably well for most<br />

of the primers we tested.<br />

6. In general, beads are concentrated by placing the 1.5-mL reaction tube on the magnetic<br />

rack for 10 to 20 s. If left for too long, the bead pellets become too tight and therefore<br />

difficult to resuspend. The supernatant is removed with a pipet tip with the tube still in<br />

the rack. Great care is taken at every step to resuspend the beads well by pipetting the<br />

suspension up and down. A drying of the bead pellets should also be avoided.<br />

7. The reagents used for sequencing (see Subheadings 2.8. and 3.8.) are available as a kit<br />

(Circumvent sequencing kit, NE <strong>Bio</strong>labs).<br />

8. Subheading 3.8., step 19 allows the separation of the labeled fragments from the beads,<br />

which facilitates gel loading. However, the presence of the beads in the sample load does<br />

not adversely affect the migration of DNA fragments.<br />

References<br />

1. Sanger, F., Nicklen, S., and Coulson, A. R. (1977) DNA sequencing with chain-termination<br />

inhibitors. Proc. Natl. Acad. Sci. USA 74, 5463–5467.<br />

2. Pfeifer, G. P., Steigerwald, S. D., Mueller, P. R., Wold, B., and Riggs, A. (1989) Genomic<br />

sequencing and methylation analysis by ligation mediated PCR. Science 246, 810–813.<br />

3. Mueller, P. R., Garrity, P. A., and Wold, B. (1993) Ligation mediated PCR for genomic<br />

sequencing and footprinting, in Current Protocols in Molecular <strong>Bio</strong>logy vol. 2 (Ausubel,<br />

F. M., Brent, R., Kingston, R. E., Moore, D. D., Seidman, J. G., Smith, J. A., et al, eds.),<br />

Current Protocols, New York, pp. 15.5.1–15.5.26.<br />

4. Quivy, J.-P. and Becker, P. B. (1994) Direct dideoxy sequencing of genomic DNA by<br />

ligation-mediated PCR. <strong>Bio</strong>techniques 16, 238–241.<br />

5. Quivy, J.-P. and Becker, P. B. (1993) An improved protocol for genomic sequencing and<br />

footprinting by ligation-mediated PCR. Nucleic Acids Res. 21, 2779–2781.<br />

6. Sambrook, J., Fritsch, E. F., and Maniatis, T. (1989) Molecular Cloning: A Laboratory<br />

Manual, Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY.<br />

7. Wu, C. (1989) Analysis of hypersensitive sites in chromatin. Methods Enzymol. 170,<br />

269–289.<br />

8. Bellard, M., Dretzen, G., Giangrande, A., and Ramain, P. (1989) Nuclease digestion of<br />

transcriptionally active chromatin. Methods Enzymol. 170, 317–346.


384 Quivy and Becker


Cloning PCR Products 385<br />

56<br />

Cloning PCR Products for Sequencing in M13 Vectors<br />

David Walsh<br />

1. Introduction<br />

Although numerous methods are now available for direct sequencing of polymerase<br />

chain reaction (PCR) products, cloning of amplified DNA for sequencing in M13 vectors<br />

remains an attractive approach because of the high quality of sequence <strong>info</strong>rmation<br />

generated from single-stranded bacteriophage DNA templates.<br />

Cloning of PCR products is in theory straightforward but in practice is often<br />

problematical, as widely reported (1–4). Difficulties are generally ascribed to modifications<br />

of the DNA termini by Taq DNA polymerase. After completion of thermal cycling,<br />

the enzyme may remain associated with DNA ends and thus interfere with subsequent<br />

ligations, unless specific steps are included for its removal or inactivation. Carryover of<br />

Taq DNA polymerase and residual dNTPs into restriction digests can also result in<br />

end filling of 5′-overhangs (5), severely reducing the efficiency of cohesive-ended<br />

cloning strategies using restriction sites within PCR primers. Removal of Taq DNA<br />

polymerase by proteinase K digestion (6) or repeated phenol/chloroform extractions<br />

(5) circumvents these problems.<br />

Furthermore, the terminal transferase activity of Taq DNA polymerase catalyzes the<br />

nontemplate-directed addition of a single nucleotide, almost invariably deoxyadenosine<br />

(dA), to the 3′ ends of amplified DNA molecules (7). The resulting “ragged ends” must<br />

be removed if blunt-end ligation to SmaI-cut vector is required. This is best achieved<br />

by utilizing the strong 3′ to 5′ exonuclease activity of T4 DNA polymerase, which<br />

in the presence of low concentrations of dNTPs removes 3′ overhangs from doublestranded<br />

DNA. Cloning of amplified DNA into linearized 5′-dephosphorylated vector<br />

also necessitates the presence of 5′ phosphate groups on the PCR products, achieved by<br />

kinasing either the primers before amplification or the PCR product itself. Conveniently,<br />

3′-dA removal by T4 DNA polymerase and 5′-phosphorylation by T4 polynucleotide<br />

kinase can be performed simultaneously (8).<br />

Difficulties in cloning PCR products as blunt-ended molecules may be avoided<br />

by incorporating restriction sites into the PCR primers and cloning products more<br />

efficiently as cohesive-ended molecules. The major problem encountered here is the<br />

failure of some restriction endonucleases to cleave toward the ends of DNA fragments.<br />

The presence of 4 bp 5′ to the recognition sequence is sufficient for efficient cutting<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

385


386 Walsh<br />

by most, but not all M13 polylinker enzymes (>75% digestion in 2 h by EcoRI, KpnI,<br />

AvaI, XmaI, PstI, BamHI, SacI, and XbaI, but 7.5 (nucleic-acidgrade)<br />

and chloroform (AR-grade). Store at 4°C in a dark glass bottle.<br />

2. Chloroform/isoamyl alcohol (241): Store at 4°C.<br />

3. Spin filtration device, such as Amicon Microcon (Beverly, MA) or Promega Wizard <br />

PCR purification unit (Madison, WI).<br />

4. Variable-speed microcentrifuge.<br />

5. TE buffer: 10 mM Tris-HCl, 1 mM EDTA, pH 8.0.<br />

6. T4 DNA polymerase.<br />

7. 10× T4 DNA polymerase buffer: 500 mM NaCl, 100 mM Tris-HCl, pH 7.9, 100 mM<br />

MgCl 2 , and 10 mM DTT, 500 µg/mL BSA.<br />

8. T4 polynucleotide kinase.<br />

9. 10× polynucleotide kinase (PNK) forward reaction buffer: 500 mM Tris-HCl, pH 7.5, 100<br />

mM MgCl 2 , 50 mM DTT, 500 µg/mL BSA.<br />

10. T4 DNA ligase.<br />

11. 10× T4 DNA ligase buffer: 500 mM Tris-HCl, pH 7.8, 100 mM MgCl 2 , 100 mM DTT,<br />

250 µg/mL BSA.<br />

12. Calf intestinal alkaline phosphatase (CIP).<br />

13. 10× CIP buffer: 200 mM Tris-HCl, pH 8.0, 10 mM MgCl 2 , 10 mM ZnCl 2 .<br />

14. Restriction enzymes at 10 U/µL and 10× reaction buffers.<br />

15. 0.5 M EDTA, pH 8.0.<br />

16. 3 M sodium acetate, pH 5.3.<br />

17. Absolute ethanol, stored at –20°C.<br />

18. Ultrapure stock of all four dNTPs: Make 20 mM stock in sterile water and store at –20°C.<br />

Avoid freeze/thaw.<br />

19. 10 mM ATP: Store in aliquots at –20°C. Avoid freeze/thaw.<br />

20. 60% (w/v) PEG 8000 in water, filter-sterilized. Store at room temperature away from<br />

direct sunlight.<br />

21. Ultrapure low-melting-point agarose.<br />

22. Glass bead DNA isolation kit, such as Geneclean (<strong>Bio</strong>101, Madison, WI).


Cloning PCR Products 387<br />

23. M13 RF DNA purchased from supplier or prepared in-house by CsCl density gradient<br />

centrifugation.<br />

24. Isopropyl-β-thiogalactopyranoside (IPTG): 20 mg/mL stock in sterile water, stored<br />

at –20°C.<br />

25. 5-Bromo-4-chloro-3-indolyl-βD-galactoside (X-gal): 20 mg/mL stock in dimethyl formamide,<br />

stored at –20°C in glass vial.<br />

26. Escherichia coli strain containing the F plasmid (e.g., JM101, JM107) and maintained<br />

on M9 minimal agar.<br />

3. Methods<br />

3.1. Purification of PCR Products from Taq DNA Polymerase, Primers,<br />

and dNTPs (see Notes 1 and 2)<br />

1. Extract the completed PCR with an equal volume of chloroform. Spin in a microfuge at<br />

13,000g for 2 min to separate the aqueous and organic phases.<br />

2. Extract the upper aqueous phase twice with an equal volume of phenolchloroform and<br />

once with an equal volume of chloroform/isoamyl alcohol.<br />

3. Transfer the upper aqueous layer (up to 100 µL) into a Microcon unit housed in a 1.5-mL<br />

Eppendorf tube and add 400 µL of TE buffer. Spin in a microfuge at 500g for 15 min<br />

(Microcon-100) or at 14,000g for 6 min (Microcon-50). Add a further 400 µL of TE to<br />

the sample reservoir, and spin as before. Volume retained will now be 50 to 100 µL. If<br />

required (PCR product present in low yield), concentration down to a volume of 10 to 20<br />

µL is achieved by a further spin cycle. Each cycle reduces the concentration of salts, PCR<br />

primers, and dNTPs by approx 95%.<br />

4. Invert the unit in a fresh tube and spin at 500g for 2 min to recover purified PCR product.<br />

5. Check recovery by agarose gel electrophoresis of an aliquot of the concentrated product.<br />

If amplified DNA is to be cloned by cohesive-end ligation via restriction sites<br />

incorporated into PCR primers, and these sites are known to cut efficiently, the purified<br />

product can now be digested with restriction endonucleases without further processing.<br />

PCR products to be cloned by blunt-end ligation or via digestion with restriction<br />

enzymes that cut inefficiently at DNA termini should be processed as follows.<br />

3.2. Simultaneous End Repair and Phosphorylation<br />

of PCR Products (see Notes 3 and 4)<br />

1. Set up a reaction containing: 100 ng to 1 µg of purified PCR product, 3 µL of 10× T4 DNA<br />

polymerase buffer, 100 µM each dNTP, 1 mM ATP, 0.5 U T4 DNA polymerase, 5 U T4<br />

polynucleotide kinase, and H 2 O to 30 µL. Incubate at 25°C for 20 min.<br />

2. Stop the reaction by incubating at 75°C for 10 min in the presence of 5 mM EDTA,<br />

pH 8.0.<br />

3. Increase the volume to 100 µL with H 2 O, and perform one extraction with phenol/<br />

chloroform and one with chloroform/isoamyl alcohol.<br />

4. Remove the aqueous phase to a fresh tube, and add 0.1 vol 3 M sodium acetate, pH 5.3, and<br />

2.5 vol cold absolute ethanol. Mix well and store at –20°C for 1 h or at –70°C for 20 min.<br />

Precipitate DNA by centrifugation at 13,000g for 10 min in a microfuge.<br />

5. Remove the supernatant carefully and add 0.5 mL cold 70% ethanol to the pellet. Spin<br />

again at 13,000g for 2 min. Discard the supernatant as before, vacuum dry the pellet<br />

(2–5 min), and finally dissolve DNA in 10 µL of TE.


388 Walsh<br />

PCR products are now flush-ended and phosphorylated at their 5′ ends, ready for<br />

direct cloning into SmaI-cut, dephosphorylated M13 vector or for concatamerization,<br />

as required.<br />

3.3. Concatamerization/Digestion of PCR Products (see Notes 5–8)<br />

1. To 10 µL PCR product, add 2 µL of 10× T4 DNA ligase buffer, 7 µL 60% PEG 8000 (20%<br />

final), and 2 U T4 DNA ligase. Incubate at room temperature overnight.<br />

2. Increase the volume to 100 µL with water, and perform one extraction with<br />

phenol:chloroform. Avoiding the white PEG precipitate at the interface, remove 10 µL<br />

of the aqueous phase to check extent of concatamerization by electrophoresis through<br />

0.8% agarose. Run out alongside an aliquot of the original PCR product and DNA size<br />

markers (see Note 8).<br />

3. If concatamerization is judged to be successful (PCR product present as trimers and<br />

larger species), extract the remaining 90 µL once with chloroform/isoamyl alcohol and<br />

precipitate with sodium acetate/ethanol as above. Dissolve in 20 µL TE.<br />

4. Add 10 U of appropriate restriction enzyme(s), 3 µL 10× reaction buffer, and H 2 O to<br />

30 µL. Incubate at the required temperature for 1 h.<br />

If cutting with two enzymes that have different salt requirements, digest with the<br />

low-salt enzyme first, heat-inactivate at 65°C for 20 min (or phenol-extract heat-stable<br />

enzymes), then adjust salt concentration with 1 M NaCl, and add 10 U of the second<br />

enzyme. Where two enzymes require completely different buffers, phenol/chloroformextract<br />

and sodium acetate/ethanol-precipitate the DNA in between each digest.<br />

5. Check to ensure the digest now contains only monomer-size PCR product by agarose gel<br />

electrophoresis. If larger species remain, indicating incomplete digestion of concatamers,<br />

add more enzyme and continue digestion.<br />

6. When digestion is complete, electrophorese the digest mixture through 0.8% low-meltingpoint<br />

agarose and recover the PCR product by glass bead isolation using Geneclean.<br />

3.4. Preparation of M13 Vector DNA for Ligation<br />

3.4.1. Digestion with Restriction Endonucleases<br />

Where digestion of the vector with two enzymes is required, the ability of each<br />

enzyme to cleave toward the end of linear DNA molecules should be considered to<br />

determine the preferred order of sequential addition New England <strong>Bio</strong>labs catalog,<br />

Reference Appendix; (see Note 9). Enzymes that cut less efficiently toward DNA<br />

termini should be used first. In this situation where directional cloning is required,<br />

digest M13 derivatives containing the polylinker in both orientations, for example,<br />

M13mp18 and M13mp19.<br />

1. Set up the following digest: 1 µg M13 RF DNA, 5 U restriction enzyme, 2 µL 10× reaction<br />

buffer, and H 2 O to 20 µL.<br />

2. Incubate for 1 h at the appropriate temperature (25°C for SmaI and 37°C for all other polylinker<br />

enzymes). Remove 1 µL to analyze extent of digestion by electrophoresis through<br />

0.8% agarose. If digestion is not complete, add more enzyme and continue incubation.<br />

3. When complete, extract the digest once with phenol:chloroform and precipitate DNA with<br />

sodium acetate/ethanol as in Subheading 3.2., steps 4–6. Dissolve the DNA pellet in 10 µL<br />

of TE and digest with a second enzyme if required.<br />

M13 DNA cut with a single enzyme should now be treated with calf intestinal<br />

alkaline phosphatase to reduce recircularization during ligation.


Cloning PCR Products 389<br />

3.4.2. Dephosphorylation of Vector DNA<br />

1. To 1 µg linearized M13 DNA in 10 µL TE, add CIP 0.05 U for 5′ overhangs, 0.5 U for<br />

3′ overhangs or blunt ends, 5 µL of 10× CIP buffer, and H 2 O to 50 µL of total volume.<br />

2. Incubate at 37°C for 60 min.<br />

3. Inactivate CIP by heating the reaction to 75°C for 10 min in the presence of 5 mM EDTA,<br />

pH 8.0.<br />

4. Extract the reaction once with phenol:chloroform and recover DNA by sodium acetate/<br />

ethanol precipitation as in Subheading 3.2., steps 4–6. Dissolve the DNA pellet in<br />

20 µL of TE.,<br />

5. Check recovery of M13 vector and insert DNA by electrophoresing an aliquot of each<br />

through 0.8% agarose.<br />

3.5. Ligation of PCR Products into M13 Vectors (see Notes 10 and 11)<br />

1. Set up the ligation reaction and add components in the following order: 50 ng of M13<br />

vector DNA, 1 µL of 10 mM ATP (1 mM final), 1 µL of 10× ligation buffer, 1– 4 µL of<br />

DNA insert (3- to 5-fold molar excess), 5 U T4 DNA ligase for blunt termini, 1 U T4 DNA<br />

ligase for cohesive termini, and H 2 O to 10 µL of total volume.<br />

2. Set up a negative control ligation in which an equal volume of water is substituted for<br />

the PCR DNA insert and a positive control ligation containing an appropriate blunt- or<br />

cohesive-ended restriction fragment, preferably of a similar size to the PCR product.<br />

3. Incubate overnight at 14°C for cohesive-end ligations or at room temperature for blunt-end<br />

ligations.<br />

4. Transform 2.5 to 5 µL of the ligation reaction into E. coli-competent cells and plate in a<br />

soft agar overlay containing 0.33 mM IPTG and 0.03% X-gal. Identify recombinant phage<br />

clones by blue/white selection (see Note 11).<br />

4. Notes<br />

1. The use of a spin filtration unit to purify the PCR product from unincorporated nucleotides<br />

and primers is only recommended if the PCR mixture does not contain unwanted species<br />

larger than the nucleotide cutoff value of the unit. For the Microcon-100, this corresponds<br />

to 300 bases (single-stranded) or 125 bp (double-stranded). If larger unwanted products are<br />

present, the target product should be purified by electrophoresis through low-melting-point<br />

agarose followed by adsorption to glass beads.<br />

2. If a spin filtration device is not available, PCR products can be partially purified from<br />

residual primers and dNTPs by precipitation with sodium acetate/ethanol, followed by<br />

a 70% ethanol wash. Better removal of such reactants can be achieved by adjusting to<br />

2 M ammonium acetate and adding 2 vol of ethanol, although it should be noted that<br />

ammonium ions are a strong inhibitor of T4 DNA polymerase and must be thoroughly<br />

removed by extensive washing in 70% ethanol before the end-repair step.<br />

3. Occasionally, PCR products end-repaired and kinased as described may fail to clone as<br />

blunt-ended molecules, most probably because of the persistence of Taq DNA polymerase<br />

bound at DNA termini. In this situation, residual enzyme can be removed by adding to<br />

the sample 50 µg/mL proteinase K in 10 mM Tris-HCl, pH 7.8, 5 mM EDTA, 0.5% (v/v)<br />

SDS, and incubating at 37°C for 30 min. Extract with phenol/chloroform, and precipitate<br />

PCR products with sodium acetate/ethanol.<br />

4. It is important not to exceed the recommended amount of T4 DNA polymerase enzyme or<br />

the incubation time of 20 to 30 min for the end-repair reaction, since both may result in<br />

excessive exonuclease activity and nonblunt “nibbled ends.” T4 DNA polymerase also has<br />

excessive exonuclease activity at higher temperatures (37°C).


390 Walsh<br />

5. For palindromic restriction enzyme sites, concatamerization of PCR products containing<br />

terminal half-sites reconstitutes the site. For example, end-to-end ligation of DNA<br />

molecules with terminal sequences GGA-3′ and 5′-TCC reconstitutes the BamHI recognition<br />

sequence. This allows the use of shorter PCR primers containing fewer extraneous<br />

nucleotides that do not hybridize to the target sequence.<br />

6. Intermolecular joining of PCR products is stimulated by macromolecular exclusion<br />

molecules, such as PEG 8000, with maximal stimulation occurring in the range 15 to<br />

25% (w/v).<br />

7. Concatamerization and digestion of PCR products containing nucleotides 5′ to the restriction<br />

site generates a small cohesive-ended fragment that can coprecipitate with the fulllength<br />

product and give rise to false positives on ligation into M13. Purification of the<br />

required product by glass bead isolation from low-melting-point agarose before cloning<br />

is therefore recommended. Eighty to ninety percent of clear plaques examined should<br />

then be found to contain the required insert. This figure falls to 10 to 20% if purification<br />

is not performed.<br />

8. To allow analysis of concatamerized PCR products by gel electrophoresis, PEG must first<br />

be removed by extracting with phenol/chloroform because the presence of the polymer<br />

prevents entry of DNA into agarose.<br />

9. When digesting M13 RF DNA with two enzymes that cut at closely spaced sites in the<br />

polylinker, it is preferable to perform the digests sequentially, rather than simultaneously,<br />

even when both enzymes are active in the same buffer, in order to maximize the cutting<br />

efficiency of each enzyme.<br />

10. The number of clear plaques that can be expected depends on the strategy being followed.<br />

Cloning via direct cutting of efficiently recognized terminal restriction sites should yield<br />

50 to 100 clear plaques per transformation. Rather fewer, typically 10 to 20, are produced<br />

by the concatamerization/digestion and blunt-end approaches.<br />

11. M13 clones containing the required insert can quickly be identified by PCR using primers<br />

used for the original amplification. Use sterile toothpicks to transfer phage particles from<br />

individual clear plaques into 0.5-mL microcentrifuge tubes containing all components of<br />

the original PCR mixture minus the template. Vortex lightly and add mineral oil. Heat<br />

to 95°C for 2 min to lyse phage and then perform 25 cycles of the original temperature<br />

regimen. Analyze by agarose gel electrophoresis.<br />

References<br />

1. Buchman, G. W., Schuster, D. M., and Rashtchian, A. (1992) Rapid and efficient cloning of<br />

PCR products using the CloneAmp system. Focus 14, 41– 45.<br />

2. Crowe, J. S., Cooper, H. J., Smith, M. A., Sims, M. J., Parker, D., and Gewert, D. (1991)<br />

Improved cloning efficiency of polymerase chain reaction (PCR) products after proteinase<br />

K digestion. Nucleic Acids Res. 19, 184.<br />

3. Krowczynska, A. M. and Henderson, M. B. (1992) Efficient purification of PCR products<br />

using ultrafiltration. <strong>Bio</strong>techniques 13, 286–289.<br />

4. Liu, Z. and Schwartz, L. M. (1992) An efficient method for blunt-end ligation of PCR<br />

products. <strong>Bio</strong>techniques 12, 28–30.<br />

5. Bennett, B. L. and Molenaar, A. J. (1994) Cloning of PCR products can be inhibited by Taq<br />

polymerase carryover. <strong>Bio</strong>Techniques 16, 32.<br />

6. Hitti, Y. S. and Bertino, A. M. (1994) Proteinase K and T4 DNA polymerase facilitate the<br />

blunt-end subcloning of PCR products. <strong>Bio</strong>techniques 16, 802.<br />

7. Clarke, J. M. (1988) Novel non-templated nucleotide addition reactions catalysed by<br />

prokaryotic and eukaryotic DNA polymerases. Nucleic Acids Res. 16, 9677–9686.


Cloning PCR Products 391<br />

8. Wang, K., Koop, B. F., and Hood, L. (1994) A simple method using T4 DNA polymerase to<br />

clone polymerase chain reaction products. <strong>Bio</strong>techniques 17, 236.<br />

9. New England <strong>Bio</strong>labs catalog (1994) Reference Appendix, pp. 180,181.<br />

10. Kaufman, D. L. and Evans, G. A. (1990) Restriction endonuclease cleavage at the termini<br />

of PCR products. <strong>Bio</strong>techniques 9, 304.<br />

11. Jung, V., Pestka, S. B., and Pestka, S. (1993) Cloning of polymerase chain reactiongenerated<br />

DNA containing terminal restriction endonuclease recognition sites, in Methods<br />

in Enzymology, vol. 218 (Wu, R., ed.), Academic, London, pp. 357–362.<br />

12. Pfeiffer, B. H. and Zimmerman, S. B. (1983) Polymer-stimulated ligation: Enhanced bluntor<br />

cohesive-end ligation of DNA or deoxyribonucleotides by T4 DNA ligase in polymer<br />

solutions. Nucleic Acids Res. 11, 7853–7871.


392 Walsh


The Vectorette Method 393<br />

57<br />

DNA Rescue by the Vectorette Method<br />

Marcia A. McAleer, Alison J. Coffey, and Ian Dunham<br />

1. Introduction<br />

A major advance in physical mapping of the human genome was the development of<br />

yeast artificial chromosome (YAC) vectors (1). This has enabled the cloning of pieces<br />

of DNA several hundred kilobases in length (2). The availability of such large cloned<br />

genomic DNA fragments means that by ordering a series of overlapping YAC clones,<br />

a contiguous stretch of DNA, several megabases in length, can be isolated around a<br />

genomic region of interest (e.g., the region of a chromosome linked to a particular<br />

disease gene). The successful isolation of terminal sequences of a given YAC can be<br />

very useful in assembling an ordered “contig” of YAC clones. Such terminal clones may<br />

be used directly as hybridization probes or sequenced and used to generate sequence<br />

tagged sites (STSs) to identify overlaps between, and isolate other, members of the<br />

contig. Several methods have been used to this end, including PCR with vector-specific<br />

primers in combination with primers designed either for repetitive elements, such as Alu<br />

sequences (3), or in combination with random nonspecific primers (4). However, these<br />

techniques rely on a suitable repetitive element or random primer sequence occurring<br />

close enough to the end of the YAC so as to be amplified by PCR. Furthermore, probes<br />

isolated in this manner may well contain highly repetitive sequences that, if unsuccessfully<br />

blocked, will increase nonspecific signal in any subsequent hybridization<br />

procedures (5).<br />

The vectorette method was originally described by Riley et al. (6). YAC DNA is<br />

digested with a restriction enzyme, and the resulting fragments are ligated to a linker<br />

molecule to create a vectorette “library,” that is, a complex mixture of restriction<br />

fragments with linker ligated to each end. Within this library are fragments that contain<br />

the YAC vector/genomic DNA junction, which includes the terminal sequences of<br />

the YAC (Fig. 1).<br />

The linker molecule consists of two long (>50 nucleotides) preannealed oligonucleotides<br />

incorporating a suitable 5′-overhang corresponding to the restriction enzyme<br />

used in the initial YAC digest. Blunt-ended linkers may also be used. Although the<br />

oligonucleotides comprising the linker are complementary at the 5′- and 3′-ends, there<br />

is a region of noncomplementarity in the middle where the two strands are unable to<br />

pair and a vectorette “bubble” is formed. The PCR is then performed on this mixture<br />

using one of two vector-specific primers (designed either for the centric or acentric<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

393


394 McAleer, Coffey, and Dunham<br />

Fig. 1. A schematic representation of the vectorette method. Solid boxes represent genomic<br />

DNA, and the hatched boxes represent YAC vector sequence. YAC DNA is digested with a<br />

restriction enzyme, X. After ligation to annealed vectorette oligos, products are amplified<br />

with a vectorette-specific primer (P2) and a primer specific for one or other of the YAC vector<br />

arms (P1). Only fragments containing vector/insert junction are amplified. Confirmation of the<br />

presence of the cloning site (CS) within the amplified fragment can be obtained by digestion<br />

of the hybrid fragment with the enzyme that cuts at the cloning site, releasing a fragment<br />

diagnostic of the vector arm (Table 2) together with one or more fragments corresponding to<br />

the genomic DNA insert. CS = cloning site.<br />

vector arms) in combination with a linker-specific primer. The linker-specific primer<br />

corresponds to the sequence of the linker ligated to the 5′-end of each DNA strand and<br />

has no complement on the other strand of the “bubble.” It is therefore unable to anneal<br />

to template until the complementary sequence has been generated by priming off the<br />

vector-specific sequence. Thus, only those fragments containing binding sites for the<br />

vector-specific primer (i.e., DNA including and immediately adjacent to the cloning<br />

site of the YAC vector) will be successfully amplified by the PCR. The amplification<br />

products may then be used as DNA probes, for DNA sequencing, or may be cloned<br />

into a suitable vector.<br />

A recent adaptation of the vectorette method has been used to isolate possible<br />

gene fragments from selected regions of the genome without prior knowledge of gene<br />

sequence (7). This method is termed Island Rescue PCR (IRP), and relies on the fact<br />

that nearly all housekeeping genes and over 40% of tissue-specific genes have a CpG<br />

island in or near the 5′-end of the gene (8). Such CpG islands have a significantly<br />

increased C + G content compared with the bulk of genomic DNA. These CpG


The Vectorette Method 395<br />

islands can be detected in native human genomic DNA, by rare-cutting restriction<br />

endonucleases that recognize unmethylated CpG-containing sequences. The principles<br />

of the vectorette method described above are used except the YAC DNA in this instance<br />

is digested with restriction endonucleases that specifically recognize CpG-containing<br />

sequences, for example, SacII, EagI. Therefore, YAC DNA will be cut at CpG-rich sites,<br />

which may be associated with a gene. The mixture is then ligated to the preannealed<br />

vectorette oligos, and PCR in this instance is driven by an Alu-specific primer together<br />

with the vectorette oligo described above. Northern blot analysis may then be used<br />

to test that amplified sequences are associated with expressed mRNAs. There are<br />

two main drawbacks to this method. First, because DNA in yeast is not differentially<br />

methylated, all CpG-containing restriction sites will be cut whether or not they are<br />

associated with an unmethylated island in native genomic DNA. Therefore, a portion<br />

of the amplified fragments may not be associated with an expressed mRNA. Second,<br />

as with all Alu-PCR based methods, there is a requirement for an Alu sequence close<br />

enough to the restriction site to allow amplification by the Taq polymerase. However,<br />

in terms of transcript mapping, where the previously described methods, for example,<br />

direct selection/cDNA enrichment (9), exon trapping (10), probing cDNA libraries<br />

directly with radiolabeled YAC DNA (11), all have limitations, IRP may prove to<br />

be a rapid and useful technique for the identification of transcriptional units within<br />

complex sources of DNA.<br />

Although the vectorette method was originally developed for rescuing the vectorinsert<br />

junctions of YACs, it may be used to isolate sequences adjacent to any known<br />

sequence, for example, the identification of intron/exon boundaries in a specified<br />

gene (12). This chapter describes in detail the application of the vectorette method to<br />

isolating terminal sequences from YACs.<br />

2. Materials<br />

All solutions should be made to the standard required for molecular biology, that is,<br />

using sterile distilled water and molecular-biology-grade reagents.<br />

1. T4 DNA ligase, 1 U/µL and 5X T4 DNA ligase buffer (0.25 M Tris-HCl, pH 7.6, 50 mM<br />

MgCl 2 , 5 mM ATP, 5 mM DTT, 25% [w/v] polyethylene glycol-8000; Gibco BRL, Paisley,<br />

Scotland).<br />

2. The sequences of the oligonucleotides used in this chapter are given in Table 1 and are<br />

taken from ref. (13). The vector-specific primers are designed against pYAC4 (these can<br />

be replaced with appropriate vector primers or Alu-specific primers if performing IRP).<br />

The vectorette oligonucleotides described are suitable for blunt-ended ligations. If desired,<br />

a suitable overhang at the 5′-end of the “top” strand oligonucleotide may be incorporated<br />

to facilitate “sticky ended” ligations.<br />

Oligonucleotides were synthesized by phosphoramidite chemistry on an Applied<br />

<strong>Bio</strong>systems 392 DNA/RNA synthesizer. After deprotection (7 h at 55°C), oligonucleotides<br />

are dried in a centrifugal evaporator (alternatively, the standard ethanol precipitation<br />

procedure may be used). Oligonucleotides used in PCR are resuspended in H 2 O to a<br />

concentration of 20 µM. Vectorette oligonucleotides are purified by high-performance<br />

liquid chromatography (12% polyacrylamide gel electrophoresis may also be used). Before<br />

use, equimolar quantities of the “top” and “bottom” oligonucleotides are preannealed in<br />

25 mM NaCl by heating at 65°C for 5 min and left to cool to room temperature. A working<br />

concentration of 1 µM is used in ligations. All oligonucleotides are stored at –20°C.


396 McAleer, Coffey, and Dunham<br />

Table 1<br />

Oligonucleotide Sequences for Vectorette PCR<br />

Vectorette oligonucleotides (for blunt-ended ligations)<br />

“Top” strand<br />

CAAGGAGAGGACGCTGTCTGTCGAAGGTAAGGAACGGACGAGAGA<br />

AGGGAGAG<br />

“Bottom” strand<br />

CTCTCCCTTCTCGAATCGTAACCGTTCGTACGAGAATCGCTGTCCTC<br />

TCCTTG<br />

Universal vectorette primer 224<br />

CGAATCGTAACCGTTCGTACGAGAATCGCT<br />

pYAC4-specific primers<br />

Centric (“left”) arm<br />

1089 CACCCGTTCTCGGAGCACTGTCCGACCGC<br />

Sup4-2<br />

GTTGGTTTAAGGCGCAAGAC<br />

pYACL (13)<br />

AATTTATCACTACGGAATTC<br />

Acentric (“right”) arm<br />

1091 ATATAGGCGCCAGCAACCGCACCTGTGGCG<br />

Sup4-3<br />

GTCGAACGCCCGATCTCAAG<br />

pYACR (13)<br />

CCGATCTCAAGATTACGGAATTC<br />

All oligonucleotide sequences are written in the 5′→3′ direction.<br />

3. PCR is performed using a GeneAmp PCR reagent kit (Perkin–Elmer, Warrington, UK)<br />

in 10 mM Tris-HCl (pH 8.3), 50 mM KCl, 1.5 mM MgCl 2 , 0.01% (w/v) gelatin containing<br />

200 µM of each dNTP and 1.0 µM of each primer. Amplitaq is added to a concentration<br />

of 1.25 U/50 µL reaction and Perfect Match (Stratagene, Cambridge, UK) to a concentration<br />

of 5 U/50 µL reaction and overlaid with mineral oil (Sigma, Poole, UK). DNA<br />

amplification is performed in an Omnigene thermocycler (Hybaid, Teddington, UK).<br />

3. Methods<br />

1. Take half an agarose plug (approx 50–100 µL containing 1–2 µg DNA) of miniprep YAC<br />

DNA (DNA in solution may also be used; see Note 1) and wash as follows: 3 × 20 min<br />

in 10 mM Tris-HCl, 0.1 mM EDTA, pH 7.4 (1 mL/plug) at 50°C. 1 × 20 min in 10 mM<br />

Tris-HCl, 0.1 mM EDTA, pH 7.4 (1 mL/plug) at room temperature.<br />

2. Preincubate plugs for 30 min at 37°C in 100 µL of the appropriate enzyme buffer (see<br />

manufacturer’s recommendation).<br />

3. Remove buffer, and replace with 100 µL of fresh enzyme buffer containing 20 to 30 U<br />

of restriction enzyme (see Note 2), and incubate overnight at the recommended temperature<br />

(usually 37°C). After digestion, the plug may be cut into three, and one portion<br />

electrophoresed through a 1.0% agarose mini gel alongside a similar amount of untreated<br />

YAC DNA to test for complete digestion. One slice may be stored dry at 4°C and redigested<br />

if incomplete digestion has occurred.<br />

4. Incubate one third of the agarose plug from step 3 in 1 mL of 1× ligation buffer for<br />

1 h on ice.<br />

5. Replace with 100 µL of fresh 1× ligation buffer. To this add 10 µL of preannealed bluntended<br />

vectorette linker (at 1 µM; see Subheading 2., item 2), that is, 10 pmol of linker.<br />

6. Heat to 65°C for 15 min to melt the agarose plug, and then equilibrate at 37°C<br />

(approx 5 min).


The Vectorette Method 397<br />

Fig. 2. Three vectorette “libraries” were created using the blunt-ended restriction enzymes:<br />

PvuII (lanes 1 and 2), StuI (lanes 3 and 4), and RsaI (lanes 5 and 6). PCR was performed using<br />

oligos specific for the centric arm of the pYAC4, 1089, and the universal vectorette oligo, 224.<br />

In A, 5 µL of Perfect Match have been added to each PCR, whereas in B, this has been omitted.<br />

Ten microliters of untreated product were loaded on a 2.5% agarose minigel in lanes 1, 3, and<br />

5, whereas samples in lanes 2, 4, and 6 were first digested with EcoRI to release the vector<br />

arm from the genomic fragment. Lane 7 contains HaeIII fragments of ΦX RF DNA (Gibco<br />

BRL, Paisley, Scotland). StuI-digested YAC has failed to produce a PCR product (A, lanes 3 and<br />

4), probably through the lack of an enzyme site close to the vector/insert junction. PvuII- and<br />

RsaI-digested YAC yields products of approximately 800 and 500 bp (lanes 1 and 5), respectively,<br />

which on digestion with EcoRI release vector fragments (V) of the predicted size 287 bp together<br />

with the terminal PvuII and RsaI fragments of the YAC insert (500 and 200 bp).<br />

7. When the reaction mix is equilibrated, add 1 µL of T4 DNA ligase (1 U/µL) and incubate<br />

at 37°C. After 1 h, add 400 µL of 10 mM Tris-HCl, 0.1 mM EDTA, pH 8.0, and mix<br />

thoroughly. The vectorette library may now be stored in aliquots at –20°C.<br />

8. Two sets of PCR mixes need to be prepared for each vectorette library constructed. The<br />

first contains a primer, 1091, specific for the “right” arm of the YAC vector (i.e., the<br />

acentric arm encoding the URA3 gene) together with the vectorette-specific oligo (224),<br />

whereas the second contains a primer, 1089, specific for the “left” arm of the YAC vector<br />

(i.e., the centric arm, which contains the CEN4 gene) together with 224 (see Table 1).<br />

Each reaction is carried out in 50 µL buffer described in Subheading 2., item 3, including<br />

5 µL of Perfect Match (see Note 3 and Fig. 2) using the following cycling conditions:<br />

94°C for 1 min, 1 cycle, followed by 93°C for 1 min, 65°C for 1 min, and 72°C for 3 min,<br />

38 cycles, and followed by 72°C for 5 min, 1 cycle. For IRP, see Note 4 for suggested<br />

primer sequences.<br />

9. Confirmation that PCR products originate from the terminal sequences of YAC clones can<br />

be obtained by demonstrating the presence of the YAC vector cloning site in the hybrid<br />

fragment. This is done by digesting the PCR product with a restriction enzyme that cleaves<br />

within the cloning site. To 9 µL of PCR product, add 1 µL of 10× restriction enzyme<br />

buffer (see manufacturer’s recommendation), 10 U of enzyme, and incubate for 1 h at<br />

37°C. When the vector is pYAC4, 10 U of EcoRI may be added directly to 9 µL of PCR<br />

product, without addition of enzyme buffer. Restriction fragments can be visualized on a<br />

2.5% agarose minigel containing ethidium bromide (0.5 µg/mL; Fig. 2). Note: ethidium<br />

bromide is a powerful mutagen and gloves should be worn at all times. The distances<br />

from the primer sequences described in Subheading 2., item 3 to the EcoRI cloning site<br />

of pYAC4 are given in Table 2.


398 McAleer, Coffey, and Dunham<br />

Table 2<br />

Positions of Primer Sequences Described in Table 1 with Respect<br />

to the EcoRI Sequence in the Cloning Site of pYAC4<br />

Centric arm<br />

Acentric arm<br />

1089→EcoRI 287 bp 1091→EcoRI 172 bp<br />

Sup4-2→EcoRI 40 bp Sup4-3→EcoRI 29 bp<br />

pYACL→EcoRI 17 bp pYACR→EcoRI 20 bp<br />

10. A second PCR may be performed to reduce the amount of vector DNA contained in the<br />

amplified product. A nested vector-specific primer that anneals closer to the cloning site<br />

(Tables 1 and 2) is used in combination with the vectorette-specific oligo. Either use<br />

1 µL of the primary PCR or toothpick the fragment found to cut with EcoRI in step 9<br />

(not the restriction digestion product) directly from the agarose gel into a PCR containing:<br />

for “left” arm products: Sup4-2 + 224 or pYACL + 224, and for “right” arm products:<br />

Sup4-3 + 224 or pYACR + 224. The same cycling conditions as those described in step 8<br />

are used, but the annealing temperature is reduced to 59°C and only 20 cycles are<br />

performed. Ten microliters may be visualized on a 2.5% agarose minigel.<br />

11. PCR products may now either be sequenced directly, radiolabeled and used as a hybridization<br />

probe (see Note 5), or subcloned using a suitable cloning system, such as pCR-Script <br />

SK (Stratagene) or TA-cloning system (Invitrogen, Leek, The Netherlands).<br />

4. Notes<br />

1. Use approx 1 µg of solution DNA for each restriction enzyme digest. Reactions should be<br />

performed in the buffers recommended by the manufacturers for 4 h at the specified<br />

temperature. Before ligation (Subheading 3., step 5), enzymes should be heat-inactivated<br />

(65°C for 15 min is usually sufficient), extracted with phenolchloroform (equal volume of<br />

ratio 11), ethanol-precipitated by standard methods (2 vol 95% ethanol with one-tenth vol<br />

3 M sodium acetate, pH 5.6) and resuspended to a concentration of 250 ng/µL. Ligations<br />

can be performed in a volume of 10 µL with 1 µL preannealed vectorette oligos.<br />

2. It is important to check that there are no recognition sites for a given restriction endonuclease<br />

between the sequences corresponding to the vector-specific primers and the<br />

cloning site. If such a site were present, it would be cleaved in the initial digest and a<br />

vector-only fragment would be amplified. Suitable enzymes for pYAC4 are RsaI, PvuII,<br />

and StuI.<br />

3. The addition of Perfect Match to the PCR reduces the number of nonspecific bands generated<br />

(compare Fig. 2A with B), although some laboratories have found little difference<br />

on its omission.<br />

4. IRP is a variant of Alu-vectorette PCR that can be used to generate probes from YACs as an<br />

alternative to Alu-PCR. For IRP, the universal vectorette primer 224 is used together with<br />

primer sequences that recognize a human Alu repeat. For example: 5′-GGATTACAGGC-<br />

GTGAGCCAC-3′ and 5′-GATCGCGCCACTGCAC TCC-3′ (both sequences taken from<br />

ref. 7). The thermocycling conditions described in Subheading 3., step 8 may also be<br />

used for these two sets of primers.<br />

5. Probes generated by this method may contain highly repetitive sequences. Therefore, it is<br />

advisable to pre-reassociate the labeled probe with total human genomic DNA prior to any<br />

hybridization procedure. Make probe up to 250 µL with H 2 O. Add 125 µL of 10 mg/mL<br />

sonicated total human DNA (Sigma) and boil for 5 min. Snap-chill on ice for 5 min, and<br />

then add probe to hybridization as normal.


The Vectorette Method 399<br />

References<br />

1. Burke, D. T., Carle, G. F., and Olson, M. V. (1987) Cloning of large segments of exogenous<br />

DNA into yeast by means of artificial chromosome vectors. Science 236, 806–812.<br />

2. Cohen, D., Chumakov, I., and Weissenbach, J. (1993) A first-generation physical map of<br />

the human genome. Nature 366, 698–701.<br />

3. Nelson, D. L., Ledbetter, S. A., Corbo, L., Victoria, M. F., Ramirez-Soli, R., Webster, T. D.,<br />

et al. (1989) Alu polymerase chain reaction: A method for rapid isolation of human-specific<br />

sequences from complex DNA sources. Proc. Natl. Acad. Sci. USA 86, 6686–6690.<br />

4. Wesley, C. S., Myers, M. P., and Young, M. W. (1994) Rapid sequential walking from<br />

termini of cosmid, P1 and YAC inserts. Nucleic Acids Res. 22, 538–539.<br />

5. Cole, C. G., Patel, K, Shipley, J., Sheer, D., Bobrow, M., Bentley, D. R., et al. (1992)<br />

Identification of region-specific yeast artificial chromosomes using pools of Alu elementmediated<br />

polymerase chain reaction probes labelled via linear amplification. Genomics<br />

14, 931–938.<br />

6. Riley, J., Ogilvie, D., Finniear, R., Jenner, D., Powell, S., Anand, R., et al. (1990) A novel,<br />

rapid method for the isolation of terminal sequences from yeast artificial chromosome<br />

(YAC) clones. Nucleic Acids Res. 18, 2887–2890.<br />

7. Valdes, J. M., Tagle, D. A., and Collins, F. S. (1994) Island rescue PCR: A rapid and<br />

efficient method for isolating transcribed sequences from yeast artificial chromosomes and<br />

cosmids. Proc. Natl. Acad. Sci. USA 91, 5377–5381.<br />

8. Larsen, F., Gundersen, G., Lopez, R., and Prydz, H. (1992) CpG islands as gene markers<br />

in the human genome. Genomics 13, 1095–1107.<br />

9. Lovett, M., Kere, J., and Hinton, L. (1991) Direct selection: A method for isolation of<br />

cDNAs encoded by large genomic regions. Proc. Natl. Acad. Sci. USA 88, 9628–9632.<br />

10. Buckler, A. J., Chang, D. D., Graw, S. L., Brook, J. D., Haber, D. A., Sharp, P. A., et al.<br />

(1991) Exon amplification: A strategy to isolate mammalian genes based on RNA splicing.<br />

Proc. Natl. Acad. Sci. USA 88, 4005– 4009.<br />

11. Elvin, P., Slynn, G., Black, D., Graham, A., Butler, R., Riley, J., et al. (1990) Isolation<br />

of cDNA clones using yeast artificial chromosome probes. Nucleic Acids Res. 18,<br />

3913–3917.<br />

12. Roberts, R. G., Coffey, A. J., Bobrow, M., and Bentley, D. R. (1993) Exon structure of the<br />

human dystrophin gene. Genomics 16, 536–538.<br />

13. Coulson, A., Kozono, Y., Lutterbach, B., Shownkeen, R., Sulston, J., and Waterston, R.<br />

(1991) YACs and the C. elegans genome. <strong>Bio</strong>essays 13, 413–417.


400 McAleer, Coffey, and Dunham


Sequencing Difficult Templates 401<br />

58<br />

Technical Notes for Sequencing Difficult Templates<br />

David Stirling<br />

1. Introduction<br />

There are a number of template types that are generally recognized as being difficult<br />

to sequence. These can include sequences with a high guanine–cytosine (G/C) content,<br />

sequences that are very rich in adenine/thymine (A/T), sequences with a marked<br />

secondary structure, or large regions of homopolymer. There is no one solution to<br />

these difficulties; however, there are a number of approaches that can be used to improve<br />

the quality of sequencing data obtained from each type of problem sequence.<br />

2. High G/C Content<br />

Sequences with high G/C content have higher melting temperatures, and the<br />

incorporation of the dye-labeled terminators is also less efficient. The addition of<br />

DMSO to a final concentration (v/v) of 5% or betaine (final concentration of 1 M ) to<br />

cycle sequencing reactions can greatly improve results. Changing the cycling conditions<br />

to use a higher melting temperature can help, as can the use of altered sequencing<br />

chemistry. Applied <strong>Bio</strong>systems produce a dGTP Big Dye kit where dITP normally<br />

used in big dye chemistry is replaced with dGTP. Adding more Taq and dNTPs can<br />

also help.<br />

3. A/T-Rich Sequences<br />

A / T-rich sequences are generally not as difficult as G/C-rich templates. A /T-rich<br />

primers will have low melting temperatures and so may need to be longer (24–26 bases)<br />

to increase the melting temperature closer to 55°C. Dye primer sequencing is generally<br />

more even through A /T regions.<br />

4. Secondary Structure<br />

Sudden loss of sequencing data generally indicates a problem with secondary<br />

structure. Where there are runs of Gs followed by runs of Cs, for instance, hairpin loops<br />

can form, impeding the progress of the polymerase and resulting in a stop in the data.<br />

Any of the approaches suggested for high G/C content sequences can be used.<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

401


402 Stirling<br />

5. Homopolymer Regions<br />

Enzyme slippage can occur when a run of the same bases occur consecutively<br />

(sometimes less than 10 for G and C). One solution is to use an anchored primer. This<br />

consists of a string of the repeated base, followed by a degenerate 3′ position of the<br />

other three bases. This allows the sequence to be resumed after the homopolymer but<br />

gives no <strong>info</strong>rmation regarding the number of bases in the repeat. Designing a primer<br />

30 to 40 bases from the start of the homopolymer and increasing the Taq concentration<br />

can improve the chances of reading through the homopolymer.


PRINS and In Situ PCR 405<br />

59<br />

PCR-Based Detection of Nucleic Acids<br />

in Chromosomes, Cells, and Tissues<br />

Technical Considerations on PRINS and In Situ PCR<br />

and Comparison with In Situ Hybridization<br />

Ernst J. M. Speel, Frans C. S. Ramaekers, and Anton H. N. Hopman<br />

1. Introduction<br />

The polymerase chain reaction (PCR) is an extremely sensitive technique allowing<br />

the detection of rare and low copy nucleic acid sequences (up to 1–10 copies in DNA<br />

or mRNA extracts from 1 million cells) by solution-phase amplification using specific<br />

primer sets and Taq DNA polymerase and visualization of the resulting PCR products<br />

by gel electrophoresis and blotting techniques (see chapters in this book). However,<br />

the obligatory cell and tissue destruction required for nucleic acid extraction does not<br />

permit the correlation of results with histopathological features or the localization<br />

of targets in specific cell types. This may now be overcome by combining recently<br />

developed micromanipulation systems, such as laser-assisted microdissection, to isolate<br />

the cells of interest (up to the level of a single cell) from a large population of cells<br />

or from the tissue, and to apply single-cell PCR on the extracted nucleic acids (for a<br />

review, see ref. 1). Alternatively, and in case it is unknown in which cell a certain target<br />

nucleic acid, for example, a virus particle, may be present, cellular localization of<br />

DNA and RNA can be accomplished by in situ hybridization (ISH). This procedure has<br />

a history of more than 30 years and has been improved continuously. In particular, the<br />

development of nonradioactive approaches and the recent implementation of tyramide<br />

signal amplification have made ISH a powerful technique for use in many applications<br />

(2). Some 10 to 15 years ago, however, ISH detection limits were only in the range of<br />

10 to 20 copies of mRNA or viral DNA per cell, and probe detection periods could be<br />

very long when using radioactive procedures (3–5). Hence, in the end of the 1980s and<br />

1990s, several strategies had been developed to improve the threshold levels as well as<br />

the efficiency of nucleic acid detection in situ, such as target and signal amplification<br />

methods (Table 1). In this chapter, we focus on the technical aspects of the primed in<br />

situ labeling (PRINS) and in situ PCR procedures, originally introduced by Koch et al.<br />

(6) and Haase et al. (7), respectively, as examples of nucleic acid target amplification<br />

methods. The pros and cons of both techniques will be discussed and compared with<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

405


406 Speel, Ramaekers, and Hopman<br />

Table 1<br />

Approaches to Amplify Nucleic Acid Target Sequences and (Immuno)<br />

Cytochemical Detection Signals in situ (Adapted from ref. 2)<br />

Target<br />

Reference<br />

Nucleic acid target amplification<br />

In situ polymerase chain reaction (in situ PCR) DNA (This chapter)<br />

Primed in situ labeling (PRINS) DNA (This chapter)<br />

and repeated /cycling PRINS<br />

Rolling circle amplification Hybridized (55)<br />

oligonucleotide<br />

(Padlock probe)<br />

In situ reverse transcriptase (RT) PCR RNA (8–10,13)<br />

In situ self-sustained sequence replication (3SR) RNA (56)<br />

In situ transcription /PRINS RNA (57,58)<br />

Detection signal amplification<br />

Branched DNA amplification (59)<br />

Catalyzed reporter deposition /tyramide signal amplification (CARD/ TSA) (2,14)<br />

Mirror image complementary antibodies (MICA) (60)<br />

Enzyme antibody polymer system (EPOS/EnVision) (61,62)<br />

Enzyme-labeled antibody–avidin conjugates (63)<br />

End Product Amplification (anti-DAB antibody strategy) (64)<br />

the current protocols for ISH using tyramide signal amplification. Special emphasis<br />

will be on the conditions needed to achieve an optimal balance between nucleic acid<br />

detection in situ and preservation of cell and tissue morphology, including discussions<br />

on sample fixation and pretreatment, oligonucleotide/probe hybridization, in situ<br />

primer extension and amplification, and detection of incorporated reporter molecules.<br />

For more comprehensive reviews and applications of PRINS, in situ PCR, and ISH,<br />

we refer to the literature (2,8–17).<br />

2. PRINS<br />

2.1. PRINS vs ISH<br />

Particularly in the field of cytogenetics, the PRINS labeling technique has become an<br />

alternative to ISH for the localization of nucleic acid sequences in chromosome and cell<br />

preparations (6,8,12,17). Occasionally, its application to the detection of chromosome<br />

copy numbers and viral DNA in frozen and formaldehyde-fixed, paraffin-embedded<br />

tissue sections have been reported (12,18,19). Whereas in an ISH approach a nucleic<br />

acid probe with incorporated reporter molecules is hybridized to its cellular target, the<br />

PRINS method is based on the use of high concentrations of unlabeled primers (restriction<br />

fragment, PCR product, or oligonucleotide) that allow a very fast hybridization<br />

(annealing) to denatured, complementary target sequences in situ (Fig. 1).<br />

These primers serve as initiation sites for in situ chain elongation catalyzed by<br />

Taq DNA polymerase (in the appropriate buffer containing 1.5 mM MgCl 2 ) using the<br />

target DNA as a template. DNA labeling takes place during the elongation step<br />

when labeled nucleotides are incorporated. Fluorochrome-labeled nucleotides can<br />

be detected directly by fluorescence microscopy, while haptenized (e.g., biotin,


PRINS and In Situ PCR 407<br />

Fig. 1. Basic differences between in ISH, PRINS, and cycling PRINS DNA labeling and<br />

direct and indirect in situ PCR. *Incorporated reporter molecule (fluorochrome or hapten [biotin,<br />

digoxigenin, dinitrophenyl]), which can be detected as described in Subheading 2.1.<br />

digoxigenin, dinitrophenyl) nucleotides can be visualized by the additional application<br />

of fluorochrome- or enzyme-conjugated avidin or antibody molecules, followed by<br />

fluorescence microscopy or brightfield visualization of enzyme reaction products<br />

(2). Image recording and analysis is usually performed by using a CCD camera or<br />

confocal scanning laser microscope. In this overview, the use of radioactivity as reporter<br />

molecule will not be considered because of disadvantages related to cost, instability,<br />

biohazard potential, the time required for autoradiographic detection, and the poor<br />

resolution of the final in situ signals.<br />

2.2. Sample Fixation and Processing<br />

In methanolacetic acid (31)-fixed metaphase spreads (adhered to noncoated glass<br />

slides), a PRINS reaction with oligonucleotides specific for centromeric or telomeric<br />

repeats usually runs for only 5 to 30 min, resulting in bright and easily localized signals<br />

and a high signal-to-noise ratio (Fig. 2A).<br />

Nevertheless, background staining might occur because of nicks in the chromosomal<br />

DNA that may act as primers for unspecific labeling. Thus, it is recommended to<br />

use freshly prepared chromosome preparations or optionally perform a DNA ligase<br />

reaction to close the nicks before PRINS labeling (17,20). Also, in interphase cells<br />

and frozen tissue sections, these repetitive target sequences can be efficiently detected<br />

provided that the specimens are adhered to coated (e.g., organosilane) glass slides,<br />

fixed in methanolacetic acid (31), and pretreated with a mild protease digestion<br />

(e.g., 100 µg/mL pepsin or 0.025% proteinase K; Fig. 2B and refs. 12,18). Because<br />

of the thorough fixation, short denaturation temperature, and speed of the reaction, the<br />

morphology of chromosomes and cell nuclei in methanolacetic acid (31)-fixed cell<br />

preparations is usually well preserved. On frozen and formaldehyde-fixed, paraffin-


408 Speel, Ramaekers, and Hopman<br />

Fig. 2. (A) Fluorescence PRINS labeling of chromosome 9 centromeres with digoxigenin/<br />

sheep anti-digoxigeninFITC in human metaphase spreads and Vectashield embedding with<br />

propidium iodide (PI) counterstaining. (B) Brightfield diaminobenzidine visualization of<br />

chromosome 9 centromeres labeled with biotin by PRINS and detected with AvidinPeroxidase<br />

in hematoxylin counterstained cell nuclei in the urothelium. (C and D) Fluorescence doublecolor<br />

PRINS labeling of chromosome 7 (C) and 9 (D) centromeres with, respectively, biotin/<br />

AvidinTexasRed and digoxigenin/sheep anti-digoxigeninFITC in human metaphase spreads and<br />

Vectashield embedding with diamidinophenylindole (DAPI) counterstaining. (E) Fluorescence<br />

ISH detection of 1 to 2 integrated HPV 16 genomes (8 kb) in the SiHa cell line using<br />

digoxigenin-labeled HPV 16 genomic DNA and rhodamin-labeled tyramide signal amplification,<br />

and Vectashield embedding with DAPI counterstaining. (F) Fluorescence PRINS labeling<br />

after pepsin pretreatment and thermal cycling in buffer without PCR reagents, of chromosome<br />

9 centromeres with digoxigenin/sheep anti-digoxigeninFITC in urothelium and Vectashield<br />

embedding with DAPI counterstaining.


PRINS and In Situ PCR 409<br />

embedded tissue sections, however, nuclear morphology may be less well preserved<br />

after PRINS DNA labeling because in our experience more powerful tissue pretreatment<br />

steps (extensive protein removal) are required than used for ISH to allow for an<br />

efficient in situ primer elongation step (18). In combination with the high denaturation<br />

temperature, as a result, discrete nuclear morphology is more difficult to maintain.<br />

Although the small size of oligonucleotides used for PRINS (usually 18 to 35 nucleotides)<br />

greatly facilitates their accessibility to genomic target sequences in comparison<br />

with the larger probes used for ISH, the subsequent in situ primer elongation reaction,<br />

thus, appears to be the success-determining step in the PRINS procedure. Particularly<br />

on tissue sections, this requires more extensive pretreatment steps than necessary for<br />

optimal ISH, which is a perhaps unexpected but important finding to realize.<br />

2.3. Primers and Hybridization Conditions<br />

The specificity of the PRINS reaction is dependent on the choice of the primer<br />

sequences as well as the conditions of primer annealing and extension used. Both single<br />

and paired primers have been described for use in PRINS. On basis of interchromosomal<br />

differences in the α-satellite and other satellite DNA repeats, human chromosomespecific<br />

PRINS primers have been designed for all human chromosomes, except for<br />

chromosome 14 and 22, which are detected simultaneously with a single 14/22 primer<br />

(21,22). Chromosome-specific PRINS primers have also been constructed for other<br />

species, such as the pig (23), as well as for other human DNA repeat regions, including<br />

telomeres and ALU sequences (24–26) and for specific single-copy genes (27,28).<br />

Optimization of the stringency of primer annealing (in sufficient reaction volume to<br />

prevent evaporation leading to concentration and temperature shifts), which reduces<br />

mispriming during the PRINS reaction, has been established by substituting the initially<br />

used thermoblocks or waterbaths by programmable thermal cyclers equipped with a<br />

flat plate block, which allow for precise and durable temperature control (up to 0.2°C<br />

accuracy). As a consequence, semi-automated PRINS protocols are now available,<br />

which offer a high reproducibility of nucleic acid labeling (8,9,29,30).<br />

2.4. Multicolor PRINS and Combination with Immunocytochemistry<br />

Multicolor detection of up to three DNA target sequences in situ have been performed<br />

using subsequent PRINS reactions with different primers and labeled oligonucleotides<br />

(30–33). This enables several targets to be analyzed simultaneously (Fig. 2C,D),<br />

which, for example, can make the evaluation of chromosome aberrations in clinical<br />

samples more robust. In addition, multicolor PRINS might be especially valuable in<br />

cases where only one or a few cells are available for analysis, for example, preimplantation<br />

genetic diagnosis. A prerequisite of performing multiple PRINS reactions in<br />

sequence is to stop the first PRINS reaction adequately and to avoid the further labeling<br />

of the first produced DNA strand with differently labeled nucleotides used in the<br />

subsequent PRINS reaction. This can be achieved by incubation with ddNTPs and<br />

Klenow DNA polymerase to block the free 3′ ends of the produced DNA strand (16,33),<br />

although others have reported that this step can be omitted, probably because of either<br />

complex chromatin conformations or incomplete DNA denaturation after relatively<br />

long extension reactions in the first PRINS reaction that hamper the access of Taq


410 Speel, Ramaekers, and Hopman<br />

polymerase (31,32). In principle, the number of in situ DNA targets to be detected<br />

simultaneously can be extended further by increasing the number of subsequent PRINS<br />

reactions applying different reporters/fluorochromes or combinations of two or more<br />

reporters in different ratios. However, the efficiency of this procedure is expected to<br />

decrease after multiple sequential PRINS reactions and, therefore, ISH is the preferred<br />

technique to use for the localization of more than 3 targets (Table 2 and refs. 34,35).<br />

PRINS has also been successfully combined with the immunocytochemical detection<br />

of proteins in multicolor approaches to, for example, immunophenotype cells harboring<br />

a specific chromosomal aberration or viral infection, to investigate chromosome<br />

distribution and segregation in cells during processes such as polyploidization and<br />

aneuploidization, and to identify possible relationships of different families of DNA<br />

sequences with, for example, proteins associated with different chromosome-specific<br />

structures, such as the kinetochore complex (see Chapter 63).<br />

Particularly, the rapidity, improved probe accessibility and lack of formamide for<br />

hybridization, thereby preventing the destruction of protein epitopes, are advantages of<br />

applying PRINS instead of ISH in these procedures.<br />

2.5. Improvement of the Detection Sensitivity<br />

The major drawback of PRINS for a long time proved to be its inability to convincingly<br />

detect single-copy gene sequences (27). This is caused by the fact that the in situ<br />

primer extension by Taq DNA polymerase in the biological material (adhered to glass<br />

slides) is limited to relatively short lengths (in the range of maximum a few hundreds<br />

of basepairs), probably caused by (1) the local chromatin organization of the target<br />

sequence; (2) the binding of (part of) the target to remnant proteins or the glass slide;<br />

and/or (3) the presence of nicks in the DNA where the polymerase reaction will stop<br />

(12). As a consequence, a single PRINS reaction to localize a single copy gene sequence<br />

with one primer (pair) will hardly result in a positive signal in the microscope, as<br />

the current detection sensitivity with ISH is approx 1 to 5 kb (2,36). The problem<br />

has now been overcome by using either multiple target-specific primers in a single<br />

PRINS reaction combined with reporter detection by the tyramide signal amplification<br />

procedure (28) or repeated PRINS reactions with the same reaction mixture and primers<br />

on specimen preparations (also called cycling PRINS) (8,20,29,37,38 and Chapter 60).<br />

Essential improvements of the first approach consisted of (1) the treatment of 1-d-old<br />

metaphase slides with 0.02 N HCl to remove loosely bound protein and thereby to<br />

render the DNA more accessible to the primer; (2) the use of multiple (four to five)<br />

primers for one locus; (3) one PRINS reaction and stringent washings in SSC to achieve<br />

optimal specificity; (4) the use of TaqStart, a monoclonal antibody against Taq DNA<br />

polymerase, which prevents nonspecific amplification and formation of primer-dimers;<br />

and (5) the use of biotin incorporation combined with biotin-labeled tyramide signal<br />

amplification (28). With this procedure a couple of single-copy genes have been<br />

detected in chromosome preparations with high efficiency (39). It will be interesting<br />

to see whether this approach can be applied to clinical cell and tissue specimens as<br />

well, for example, for rapid and reliable detection of microorganisms and chromosomal<br />

alterations in cancer specimens. Alternatively, cycling PRINS has been used on<br />

metaphase spreads and blood smears to detect low and single-copy DNA sequences<br />

resulting in more intense (up to 15×) fluorescence in situ signals than seen after a single


PRINS and In Situ PCR 411<br />

PRINS reaction (Fig. 1, refs. 20,29,37,38,40). The repeated PRINS reactions appear<br />

to label not only the target sequence (as observed after the last PRINS cycle) but also<br />

the DNA that has been synthesized during the preceding cycles. Several protocols have<br />

been described with suggestions for optimal gene localization, including (1) the use<br />

of single or multiple primers generating DNA products of 250 to 550 bp in length; (2)<br />

alternative fixation and pretreatment protocols (e.g., ethanol and microwave treatment<br />

instead of methanol:acetic acid [31] fixation); 3) the use of a hot start (applying the<br />

reaction mixture and/or Taq polymerase at the annealing or a higher temperature to<br />

the slide to avoid mispriming or primer oligomerization during PRINS/PCR) before 1<br />

to 20 PRINS cycles, eventually preceding by PCR with the same primers but without<br />

labeled nucleotides; and (4) adaptation of the PRINS reaction volume and conditions<br />

on the glass slide during cycling. However, despite these recommendations, a main<br />

concern comprises the reproducibility of cycling PRINS, because in most studies<br />

variable and relatively low frequencies (10–70%) of chromosomes or cells harboring<br />

PRINS signals were reported. Besides the discussed possibilities that may result in<br />

background (nicks, mispriming), the major disadvantage of applying multiple cycles<br />

of PRINS (or PCR, see later) on cell and tissue preparations is the inevitable diffusion<br />

of newly synthesized DNA products, inextricably bound up with the denaturation<br />

steps, from the site of synthesis inside and/or outside the cells, followed by possible<br />

extracellular generation of amplificants (Fig. 1. refs. 9,20,29,41–43). Several studies<br />

have provided evidence for this phenomenon by demonstrating the expected DNA<br />

products using gel electrophoresis of the reaction mixture after cycling PRINS. The<br />

observation of stronger but often more diffuse and less discretely localized PRINS<br />

signals in only a low percentage of chromosomes or cell nuclei fits in with the<br />

view that labeled DNA in these cases is retained at or near the site of synthesis by<br />

possible entrapment in the chromatin or nonspecific binding to the surrounding cellular<br />

structures, whereas diffusion of DNA away from the site of synthesis has been taken<br />

place in the remaining negative chromosomes or nuclei. That diffusion occurs may be<br />

explained by the fact that methanolacetic acid (31) fixation is an effective procedure<br />

to extract proteins, thus limiting the possibillities to entrap synthesized DNA molecules<br />

efficiently in the chromosome structures. Although postfixation in paraformaldehyde<br />

after cycling PRINS has been suggested to reduce diffusion of produced DNA (17), this<br />

will of course only be of help in the chromosomes and nuclei harboring signals.<br />

Thus, because of several drawbacks concerning the efficiency and reproducibility of<br />

cycling PRINS in localizing either low or single-copy DNA sequences in situ, a single<br />

PRINS reaction with multiple primers combined with tyramide signal amplification<br />

is recommended for this purpose (as a rapid alternative to ISH with locus-specific<br />

probes).<br />

3. In Situ PCR<br />

3.1. In Situ PCR vs Cycling PRINS<br />

In line with the development of PRINS, several groups working in the fields of<br />

pathology and microbiology have introduced the successful combination of PCR and<br />

ISH to visualize specific amplified nucleic acid sequences in cell and particularly<br />

formaldehyde-fixed, paraffin-embedded tissue preparations. Most studies have focused<br />

on the detection of (pro)viral (foreign) nucleic acid sequences, but in addition the


Table 2<br />

Comparison of PRINS, in Situ PCR, and ISH for Localization of DNA Target Sequences in Situ<br />

PRINS In situ PCR ISH<br />

Main application Chromosome and gene identification and Detection of altered genes or foreign As for PRINS and in situ PCR, and<br />

quantification in cells and chromosomes (e.g., viral) DNA in paraffin-embedded localization of DNA sequences to study<br />

for use in clinical and cancer genetics and tissue section for use in molecular 3D organization of the interphase nucleus<br />

preimplantation genetic diagnosis pathology<br />

Crucial steps Primer diffusion to DNA target Primer diffusion to DNA target Probe diffusion to DNA target<br />

(see Fig. 1) Primer hybridization Primer hybridization Probe hybridization<br />

Primer elongation and DNA labeling<br />

Primer elongation (and DNA labeling)<br />

by Taq polymerase by Taq polymerase<br />

PCR while limiting diffusion of newly<br />

synthesized DNA products<br />

Procedure<br />

Specimen, fixative, and Ch: Methanolacetic acid (31) Ce, Ti: 4–10% (para)formaldehyde Ch, Ce: Methanolacetic acid (31),<br />

preferred protein removal Ce, Ti: Methanolacetic acid (31) and tuned proteinase K digestion to allow Ce: 70% ethanol and mild pepsin<br />

procedure and mild pepsin digestion reagents to access the target DNA Ti: 4–10% (para)formaldehyde combined<br />

(Ch = chromosomes, but to limit diffusion of PCR products with strong pepsin and 1M sodium<br />

Ce = cells, thiocyanate or microwave<br />

Ti = tissues)<br />

Main probe type Single or multiple Oligonucleotide primer pair Cloned genomic or cDNA (1–100 kb) in<br />

Oligonucleotide(s) (20–30 mer) (20–30 mer) plasmid, cosmid, PAC or BAC vector<br />

DNA labeling in situ Specimen denaturation Incorporation of labeled Hybridization of a denatured, labeled<br />

Primer annealing nucleotides during PCR (10–30 cycles) DNA probe on the pretreated, denatured<br />

Incorporation of labeled on the pretreated specimen specimen<br />

nucleotides by DNA polymerase (= direct in situ PCR), or PCR followed<br />

by FISH detection of amplified DNA<br />

(= indirect in situ PCR)<br />

Specificity Primer annealing Primer annealing and/or ISH after PCR Probe hybridization and stringent washes<br />

Label detection Direct (fluorochromes) or Idem Idem<br />

indirect (antibody detection<br />

and signal amplification)<br />

Advantages Good accessibility of reagents by High sensitivity (target few hundred bp) Good accessibility of reagents by<br />

optimal protein removal optimal protein removal<br />

412 Speel, Ramaekers, and Hopman


Direct labeling during 1 PCR cycle<br />

Relative high sensitivity (target 1–5 kb)<br />

Multiple-target detection (up to 3)<br />

High specificity and efficiency<br />

Use of primers with different<br />

Accurate DNA localization<br />

annealing temperatures<br />

Well-preserved morphology<br />

Relative accurate DNA localization Multi-target detection (up to 32)<br />

Short turnaround time (few hours)<br />

Disadvantages Limited primer elongation due to nicks Only reasonable accessibility of PCR Relative long turnaround time (1–2 days)<br />

and target DNA bound to the glass slide reagents and poor primer elongation due<br />

Relative low sensitivity (single-copy DNA) to suboptimal protein removal at the start<br />

Relative poor morphology on tissue<br />

Increasing number of PCR cycles improves<br />

Subsequent PRINS reactions needed primer elongation but also diffusion of PCR<br />

for multiple-target PRINS products. This results in:<br />

Nonspecific incorporation of labeled<br />

Low amplification efficiency, variable<br />

nucleotides (nicks, mispriming) leading reproducibility and poor DNA localization<br />

to limited signal-to-noise ratio<br />

Relative poor morphology due to heating<br />

steps in the PCR procedure<br />

Further disadvantages:<br />

Nonspecific incorporation of labeled<br />

nucleotides at nicks (direct in situ PCR) and<br />

amplification due to mispriming<br />

Only single-target detection<br />

Relative long turnaround time (1–2 days)<br />

Comments Muliple oligonuclotides and signal Signal amplification may improve Signal amplification have improved<br />

amplification have improved signal-to- signal-to-noise ratio signal-to-noise ratio and evaluation<br />

noise ratio and enabled single copy of results<br />

gene detection<br />

PRINS and In Situ PCR 413


414 Speel, Ramaekers, and Hopman<br />

technique has also been applied to identify endogenous DNA and RNA sequences in<br />

human cells (for reviews, see refs. 8–10,13). In situ PCR techniques are theoretically<br />

straightforward and comprise (1) sample fixation and pretreatment to improve the<br />

accessibility of the target nucleic acid sequences by the PCR primers, nucleotides, and<br />

Taq polymerase enzymes and to avoid diffusion of PCR-generated amplificants; (2)<br />

PCR amplification in the cell by Taq polymerase using the target DNA as a template; and<br />

(3) direct (by incorporation of labeled nucleotides during PCR) or indirect (by ISH<br />

with labeled nucleic acid probes) detection of the amplified nucleic acid molecules.<br />

Visualization of labeled target or probe DNA can be performed as described in<br />

Subheading 2.1. Thus, in principle direct in situ PCR is identical to the cycling PRINS<br />

procedure described above when cellular DNA is directly labeled during PCR cycling<br />

(Fig. 1). Consequently, the conditions for optimal chromosome labeling by (cycling)<br />

PRINS will also apply for the direct in situ PCR approach, although they need to be<br />

adapted for the specimen of interest and the fixative used to process cells and tissues.<br />

3.2. Sample Fixation and Processing<br />

It has been reported that 4 to 10% buffered formaldehyde is the fixative of choice for<br />

successful in situ PCR on cells and (frozen) tissues and that these specimens should be<br />

adhered to coated (e.g., organosilane) glass slides to prevent loss of tissue adherence<br />

during the in situ PCR procedure (5,9). In addition, cell and tissue preparations need<br />

to be subjected to a protease (pepsin, trypsin, or protease K) treatment to create holes<br />

in cellular membranes and remove cross-linked DNA-binding proteins from nuclear<br />

DNA, thereby facilitating the accessibility of the nucleic acids in situ by the PCR and<br />

detection reagents, as well as the ISH probes (in the indirect in situ PCR procedure).<br />

Importantly, for every protease and tissue, the optimal balance between time and<br />

concentration should be determined (usually 5 to 30 min of 2 mg/mL protease at room<br />

temperature or 37°C) to avoid overdigestion, leading to poor tissue morphology and<br />

possible leakage (diffusion) of amplified products from the cell in which they were<br />

generated, or insufficient protein removal, resulting in a decreased or completely absent<br />

in situ signal caused by very inefficient or failure of amplification (5,9). A striking<br />

point to notice here is that the generally used protease pretreatment for in situ PCR<br />

is less powerful than the one applied usually for optimal ISH on formaldehyde-fixed,<br />

paraffin-embedded preparations, in which prior to the protease digestion treatments with<br />

1 M sodium thiocyanate and, optionally, 85% formic acid/0.3% H 2 O 2 are performed<br />

to achieve optimal conditions for ISH while preserving nuclear morphology (44).<br />

Although this may still be sufficient to allow the PCR reagents to reach the target DNA,<br />

we know from our own results that efficient in situ primer extension rather requires a<br />

more powerful tissue pretreatment step than used for ISH (see Subheading 2.2. and<br />

3.3.2.). Indeed, we have not been able to detect human centromere repeats by PRINS<br />

labeling using tissue pretreatment conditions optimal for ISH with centromere-specific<br />

plasmid probes (unpublished results), indicating that in situ primer extension during the<br />

first couple of cycles of PCR on the slide is most likely impossible or very inefficient.<br />

This might in part explain the low amplification efficiency (restricted sensitivity)<br />

often obtained by in situ PCR, which furthermore can be caused by diffusion of the<br />

synthesized DNA amplificants outside the cell (see Fig. 1 and below).


PRINS and In Situ PCR 415<br />

3.3. In Situ Amplification and Detection of Intracellular PCR Products<br />

PCR protocols with optimal stringency of primer annealing and using generally<br />

longer extension times, increasing concentrations of Taq polymerase and MgCl 2 , and /or<br />

the addition of bovine serum albumin in the reaction mixture on the cell and tissue<br />

preparations as compared with solution-phase PCR have been established by using<br />

the newly designed programmable thermal cyclers as described for (cycling) PRINS<br />

(see Subheading 2.3. and refs. 5,9,30). After PCR amplification, visualization of<br />

intracellular PCR products is achieved either directly through immunohistochemical<br />

detection of labeled nucleotides (see Subheading 2.1.) that have been incorporated<br />

into PCR products during thermal cycling (direct in situ PCR) or indirectly by ISH<br />

with a labeled probe and subsequent immunohistochemical detection (see Chapter<br />

63 and refs. 5,9).<br />

3.3.1 Direct In Situ PCR<br />

Although direct in situ PCR is more rapid than indirect in situ PCR by eliminating<br />

the need for subsequent ISH, this procedure has proved unreliable with respect to<br />

the specificity of the results obtained (9,43,45). Even when the hot start procedure<br />

is performed (see Subheading 2.4.), the direct detection approach yields significant<br />

false-positive results, especially when working with tissue sections that have been dried<br />

at 56 to 65°C for several hours (introduction of nicks in the DNA) (5,9). This is the<br />

result of a number of artifacts, including incorporation of labeled nucleotides into (1)<br />

nonspecific PCR products resulting from mispriming (“endogenous priming” artifacts)<br />

and primer oligomerization, which seems to occur less likely inside nuclei (5) but may<br />

play a role during extracellular amplification of diffused DNA products); (2) singleand<br />

double-stranded nicks introduced by tissue fixation, cutting, and drying; and (3)<br />

fragmented DNA undergoing “repair” by DNA polymerase (“repair” artifacts). Repair<br />

artifacts may particularly be evident in apoptotic cells or samples that have been<br />

pretreated with DNase before in situ reverse transcriptase PCR for mRNA detection<br />

(9,45). These artifacts may only slightly be reduced by using an exonuclease-free<br />

DNA polymerase or by carrying out a DNA ligase reaction to close the nicks or a<br />

ddNTP reaction to prevent a subsequent extension reaction (9,17,33). Thus, caution<br />

and adequate use of appropriate controls are recommended in the interpretation of data<br />

produced by direct in situ PCR (Table 3 and Subheading 3.4.).<br />

3.3.2. Indirect In Situ PCR<br />

The indirect in situ PCR technique, therefore, is the preferred approach to use,<br />

because the intracellular target-specific PCR products are identified by ISH, whereas<br />

the nonspecific amplificants will not be detected. Probes targeted to regions in between<br />

the primers used for PCR, usually oligonucleotide probes of 20 to 40 bases, represent<br />

the ideal ISH probes for reasons of specificity because they assure that the detected<br />

signal is the PCR product and not the result of for example primer oligomerization. ISH<br />

with these small probes, however, is usually less sensitive than with cDNA or genomic<br />

probes. Because primer oligomerization seems not to occur inside nuclei during PCR<br />

(10), the latter probes are recommended because of the increased number of reporter<br />

molecules they carry, leading to a higher detection sensitivity. Immunohistochemical


Table 3<br />

Summary of Control Experiments Required for DNA Detection by PRINS, in situ PCR, and ISH (Modified from ref. 9).<br />

Method Control experiment Purpose<br />

General Use of known positive and negative control samples Control for specificity and sensitivity<br />

as well as mixtures of these (cells) in different ratios<br />

Immunophenotpying the cell types of interest<br />

Control for specificity and sensitivity<br />

Solution-phase PCR on extracted DNA of sample under<br />

Control for sensitivity, false negative results and DNA quality<br />

study or directly on mildly fixed cell suspension<br />

Omission of primary antibody in immunohistochemical<br />

Control for background induced by endogenous enzyme activity<br />

detection (colorimetric detection) or nonspecific sticking of secondary<br />

and/or tertiary detection reagents to sample and/or glass slide<br />

Use of no, random, or irrelevant (labeled) primers/probes,<br />

Control for background induced by endogenous enzyme activity<br />

or vector sequences without a target-specific insert (colorimetric detection) or non-specific sticking of probe<br />

and/or detection reagents to sample and/or glass slide<br />

(Cycling) PRINS Omission of DNA polymerase Control for background induced by endogenous enzyme activity<br />

and (colorimetric detection) or nonspecific sticking of detection<br />

Direct in situ PCR reagents to sample and/or glass slide<br />

Omission of primers<br />

Control for artifacts related to endogenous priming, DNA repair<br />

and extension of nicks present in the DNA<br />

Indirect in situ PCR Omission of DNA polymerase Control for effect of amplification and sensitivity of ISH<br />

416 Speel, Ramaekers, and Hopman


PRINS and In Situ PCR 417<br />

detection of labeled nucleotides within target or probe DNA can be performed as<br />

described previously (Subheading 2.1. and refs. 2,46). Because of the autofluorescence<br />

often present in tissue preparations, the limited stability of fluorochromes, and the<br />

preference of histopathologists to analyze permanent preparations by brightfield<br />

microscopy, colorimetric detection systems have been most frequently used (5,9). In<br />

these procedures, the activity of alkaline phosphatase or horseradish peroxidase, coupled<br />

to (immuno)histochemical reagents (e.g., antibodies or (strept)avidin molecules), is<br />

visualized by enzyme precipitation reactions (NBT/BCIP and DAB are most often used,<br />

respectively). Strikingly, in most cases suboptimal detection formats have been used,<br />

combining, for example, only a single antibody layer (alkaline phosphatase-conjugated<br />

anti-digoxigenin Fab fragments) with a relatively poor localizing enzyme reaction<br />

(NBT/BCIP) and omitting the possibility to apply the sensitive tyramide signal<br />

amplification procedure (2,14). As a consequence, relatively poor signal-to-noise ratios<br />

have often been obtained, which may further contribute to the restricted increase in<br />

detection sensitivity obtained by (in)direct in situ PCR when compared with ISH.<br />

Moreover, this might also explain in part the relatively low sensitivity reported with<br />

the ISH procedure when the PCR step is omitted, resulting in a detection limit of only<br />

20 to 40 viral copies per cell (4,5). As can be seen in Fig. 2E, current optimal ISH<br />

protocols (44) are able to identify the 1 to 2 copies of HPV 16 DNA integrated in the<br />

genome in the often used SiHa control cell line without any prior PCR step.<br />

Because the combination of PCR and ISH in the indirect in situ PCR method<br />

is essential to guarantee that the obtained signal is specific, one may consider this<br />

approach also as a rather cumbersome ISH method, in which sample pretreatment<br />

consists of fixation and protease digestion in combination with heating (thermal<br />

cycling) during nucleic acid amplification (by PCR). Moreover, the increase in detection<br />

sensitivity as compared with conventional ISH is rather limited even after optimization<br />

of the procedures, and cell morphology and nucleic acid localization are often poor<br />

(see Subheading 3.4. and Table 2). Furthermore, the question arises if the increase<br />

in the final ISH signal intensity is really caused by the in situ amplification of target<br />

sequences alone, or is rather the result of heating the specimen by thermal cycling,<br />

because it has been shown that microwaving or thermal cycling without in situ PCR can<br />

also result in increased sensitivity of ISH (47,48). In this respect, we have reported a<br />

similar effect localizing human centromere 9-specific DNA in routinely-fixed, paraffinembedded<br />

tissue sections, where only after pepsin digestion in combination with<br />

thermal cycling in buffer without PCR reagents it proved to be possible to perform a<br />

primed in situ labeling reaction with a chromosome 9-specific primer with positive<br />

outcome (Fig. 2F and ref. 14).<br />

Indirect in situ PCR protocols are available for the simultaneous detection of two<br />

DNA targets in formaldehyde-fixed, paraffin-embedded tissue sections as well as<br />

for the detection of DNA in combination with a protein detected subsequently by<br />

immunohistochemistry (49). However, without using appropriate controls (Table 3)<br />

to adequately address the possible leakage of synthesized DNA molecules to targetnegative<br />

sites, as is the principal drawback of indirect in situ PCR, caution is recommended<br />

by interpretation of the data.


418 Speel, Ramaekers, and Hopman<br />

3.4 Sensitivity and Evaluation of In Situ PCR Results<br />

Only a few studies have tried to give an indication of the increase in in situ signal<br />

intensity by comparing indirect in situ PCR with ISH. In the most optimal situation,<br />

a 50-fold increase in sensitivity has been reported when cell preparations were used<br />

(7,50). This relatively poor result after PCR amplification thus lies far beneath the<br />

amplification efficiency that can be achieved by solution-phase PCR. As has been<br />

discussed above, this is the result of many factors, including suboptimal sample<br />

pretreatment, inefficient and nonspecific in situ primer extension, as well as diffusion of<br />

PCR products because of the denaturation steps during PCR. In addition, background<br />

signals introduced by the ISH procedure, as a result of nonspecific binding of probe and<br />

detection reagents to the sample and the coated glass slide as the result of suboptimal<br />

stringent washings after probe hybridization and application of too less diluted probe<br />

and/or detection conjugates, may even further contribute to a decrease of the difference<br />

in signal intensity between in situ PCR and ISH. Because so many factors may influence<br />

the final signal intensity observed, it is not advisable to use in situ PCR results for<br />

quantification purposes.<br />

Nevertheless, an amplification factor of 10- to 50-fold would still be a very acceptable<br />

increase in sensitivity, provided that reproducible and specific results would be<br />

obtained by indirect in situ PCR. However, despite the fact that the detection of one<br />

target copy per cell have been reported by using a single primer pair that amplifies<br />

a sequence of a few hundred basepairs (5,10), indirect in situ PCR appears also to<br />

be hampered by restricted specificity of results and, moreover, is unable to distinctly<br />

localize nucleic acid sequences in cell and tissue preparations (9,51). Even when<br />

assuming that PCR conditions are optimal and artifacts caused by mispriming and<br />

nicks are eliminated, the diffusion of generated PCR products from the site of synthesis<br />

inside and/or outside the cell is an almost impossible process to control. Therefore,<br />

many creative approaches intended to minimize the impact of diffusion have been<br />

described, such as optimal sample fixation and pretreatment, reduction of PCR cycle<br />

numbers, generation of longer or more complex PCR products, incorporation of labeled<br />

nucleotides to make bulkier PCR products, and the embedding of samples in agarose<br />

or protein matrices (9,43,50,52). Nevertheless, in most cases only a discrimination<br />

between positive and negative nuclei can be determined, whereas for example the<br />

number and site of viral integration in the nucleus, as shown in Fig. 2E by ISH,<br />

as well as the discrimination between viral integration and replication, are almost<br />

impossible to identify.<br />

Thus, appropriate controls at each step in the (in)direct in situ PCR procedure are<br />

essential to demonstrate specificity and to correctly interpret the results (Table 3 and<br />

refs. 5,9,42). Control experiments should include (1) samples that either harbor or<br />

lack the target of interest (or a known mixture of these cells to verify the expected<br />

result) or use of irrelevant primers that cannot find targets in the cells under study (e.g.,<br />

viral-specific primers in uninfected cells); (2) omission of the DNA polymerase from<br />

the PCR mixture to detect nonspecific sticking of ISH probes and detection reagents<br />

to the slide; (3) omission of primers to detect artifacts related to endogenous priming<br />

(nicks in the DNA) and DNA repair in the direct in situ PCR approaches. In case<br />

of reverse transcription in situ PCR to detect (m)RNA sequences in situ, RNase


PRINS and In Situ PCR 419<br />

pretreatment of samples and the omission of the reverse transcription step are important<br />

controls.<br />

4. Conclusion<br />

During the past 10 to 15 yr, several strategies have been developed to improve the<br />

threshold levels of nucleic acid detection in situ by either labeling and/or amplification<br />

of target nucleic acid sequences (prior to ISH), for example, (cycling) PRINS and in<br />

situ PCR, or by amplification of the detection signals after the hybridization procedure,<br />

for example, PRINS or ISH followed by tyramide signal amplification (Table 1).<br />

Although PRINS and in situ PCR have been developed and applied in different research<br />

disciplines and only sporadically have been compared with each other and/or with ISH<br />

in the same study, all three techniques have been extensively studied and optimized<br />

to cope with the central question how to achieve the most optimal balance between<br />

specific in situ nucleic acid detection with a high signal-to-noise ratio and preservation<br />

of cell and tissue morphology. As a result, all techniques are capable to detect repetitive<br />

as well as single copy DNA sequences to date. However, because of evident differences<br />

in specificity, reproducibility, and detection sensitivity between PRINS, in situ PCR,<br />

and ISH, in our opinion the most suitable technique to be used for nucleic acid detection<br />

in biological samples is ISH, optionally combined with tyramide signal amplification<br />

to achieve the highest sensitivity.<br />

In methanol:acetic acid (31)-fixed chromosome and cell preparations, a single<br />

PRINS reaction is a suitable alternative to ISH allowing for the rapid, specific, and<br />

reliable localization of repetitive and single copy DNA sequences (28,39). In the latter<br />

case, PRINS should be conducted with multiple carefully selected and target-specific<br />

oligonucleotide primers and be combined with tyramide signal amplification. The exact<br />

detection limit, however, is unknown, and will be dependent on the ratio between the<br />

obtained specific PRINS signal and the background noise resulting from mispriming<br />

and labeling of nicks in the DNA, for which control experiments should be performed.<br />

Cycling PRINS is not recommended because until now this technique has not proved<br />

to be reliable with respect to efficiency and specificity because of problems of DNA<br />

diffusion and nonspecific labeling of DNA (29,38,40,51). Although initially the turnaround<br />

time (usually 16 h) and the generation and expensive purchase of probes have<br />

been considered as real disadvantages of ISH as compared with PRINS (12), the current<br />

situation is clearly a different one. ISH can be performed in very short turnaround times<br />

if repetitive and directly labeled probes are used. Furthermore, ISH probes are well<br />

available now because of the complete sequencing of the human and other genomes,<br />

and although still expensive can be used for the specific localization of DNA probes<br />

even under 1 kb (36) as well as for the simultaneous localization of 24 or more DNA<br />

targets in chromosomes and cells (34,35). In contrast, in a single PRINS reaction a<br />

huge amount of primer(s) is used together with Taq DNA polymerase and the amount<br />

of hapten-labeled nucleotides that is normally used for labeling of 1 to 2 µg of ISH<br />

probe. In addition, the use of tyramide signal amplification required for the localization<br />

of single copy genes will lead to high costs if applied for clinical diagnostics as a<br />

result of the fact that it is a patented trademark. Moreover, only a limited number of<br />

targets can be detected simultaneously by subsequent PRINS reactions, and additional


420 Speel, Ramaekers, and Hopman<br />

advantages of PRINS over ISH, for example, the ability to discriminate between human<br />

centromeres 13 and 21 for chromosome enumeration analysis in prenatal diagnosis,<br />

have been counteracted by the development of specific probe sets that combine<br />

chromosome-specific single-copy or chromosome painting probes for discrimination.<br />

In methanolacetic acid (31)-fixed and protease-pretreated frozen tissue sections<br />

and cell preparations (touch preparations, blood smears, smears of bone marrow<br />

aspirates) so far only repetitive sequences have been visualized efficiently by a single<br />

PRINS reaction (8,12,18). Single-copy DNA sequences have been successfully detected<br />

by cycling PRINS and ISH (2,40), but because of the drawbacks involved with the<br />

use of the cycling PRINS procedure as described above an ISH approach is preferred<br />

for this purpose. The most optimal results have been obtained by applying (single<br />

or multiple-target) ISH to 70% ethanol-fixed samples that have been pretreated with<br />

pepsin (44,53).<br />

A couple of studies have compared DNA detection, particularly genomic human<br />

papillomavirus DNA, in formaldehyde-fixed, paraffin-embedded cell and tissue preparations<br />

by in situ PCR, PRINS, and ISH (5,19,51,54). With all techniques HPV-specific<br />

sequences were efficiently detected, but only the one or two HPV 16 DNA sequences,<br />

known to be integrated in the genome of the human SiHa cell line, could be visualized by<br />

ISH or indirect in situ PCR (see also Fig. 2E). Interestingly, two of these studies (51,54)<br />

reported the use of ISH in combination with tyramide signal amplification as the method<br />

of choice, since it provided the same sensitivity and a much better reproducibility and<br />

reliability than the more cumbersome and poorly reproducible indirect in situ PCR<br />

method. Apparently, under optimal conditions ISH combined with the high amplification<br />

power of the signal amplification system can result in the same detection limits as<br />

can be reached by the relatively low amplification efficiency of indirect in situ PCR<br />

(as compared with solution-phase PCR). Furthermore, it provides optimal localization<br />

of signals and discrimination between viral integration and replication within wellpreserved<br />

cells and nuclei (Hopman et al., manuscript in preparation). This is almost<br />

impossible to achieve with in situ PCR because of diffusion of PCR products in the cell<br />

nucleus as well as leakage out of the nucleus to target-negative cells, which is still the<br />

most important disadvantage of an in situ PCR approach.<br />

References<br />

1. Hahn, S., Zhong, X. Y., Troeger, C., Burgemeister, R., Gloning, K., and Holzgreve, W.<br />

(2000) Current applications of single-cell PCR. Cell. Mol. Life Sci. 57, 96–105.<br />

2. Speel, E. J. M. (1999) Detection and amplification systems for sensitive, multiple-target<br />

DNA and RNA in situ hybridization: looking inside cells with a spectrum of colors.<br />

Histochem. Cell <strong>Bio</strong>l. 112, 89–113.<br />

3. Höfler, H., Childers, H., Montminy, M. R., Lechan, R. M., Goodmann, R. H., and Wolfe,<br />

H. J. (1986) In situ hybridization methods for the detection of somatostatin mRNA in tissue<br />

sections using antisense RNA probes. Histochem J. 18, 597–604.<br />

4. Burns, J., Graham, A. K., Frank, C., Fleming, K. A., Evans, M. F., and McGee, J. O’. D.<br />

(1987) Detection of low copy human papillomavirus DNA and mRNA in routine paraffin<br />

sections of cervix by non-isotopic in situ hybridization. J. Clin. Pathol. 40, 858–864.<br />

5. Nuovo, G. J. (2000) In situ localization of PCR-amplified DNA and cDNA, in In situ<br />

hybridization protocols. Methods in Molecular <strong>Bio</strong>logy (Darby, I. A., ed.), Humana Press,<br />

Totowa, NJ, pp. 217–238.


PRINS and In Situ PCR 421<br />

6. Koch, J., Kölvraa, S., Petersen, K. B., Gregersen, N., and Bolund, L. (1989) Oligonucleotidepriming<br />

methods for the chromosome-specific labelling of alpha satellite DNA in situ.<br />

Chromosoma 98, 259–265.<br />

7. Haase, A. T., Retzel, E. F., and Staskus, K. A. (1990) Amplification and detection of<br />

lentiviral DNA inside cells. Proc. Natl. Acad. Sci. USA 87, 4971– 4975.<br />

8. Gosden, J. R., ed. (1997) PRINS and In Situ PCR protocols. Methods in Molecular <strong>Bio</strong>logy,<br />

Vol. 71, Humana Press, Totowa, NJ.<br />

9. Long, A. A. and Komminoth, P. (1997) In situ PCR, in PRINS and in situ PCR protocols.<br />

Methods in Molecular <strong>Bio</strong>logy, Vol. 71 (Gosden, J. R., ed.), Humana Press, Totowa, NJ,<br />

pp. 141–161.<br />

10. Nuovo, G. J. (1997) PCR in situ hybridization, protocols and applications, 3rd ed.,<br />

Lippincott-Raven, New York.<br />

11. Totos, G., Tbakhi, A., Hauser-Kronberger, C., and Tubbs, R. R. (1997) Catalyzed reporter<br />

deposition: a new era in molecular and immunomorphology—Nanogold-silver staining and<br />

colorimetric detection and protocols. Cell Vision 4, 433– 442.<br />

12. Wilkens, L., Tchinda, J., Komminoth, P., and Werner, M. (1997) Single- and double-color<br />

oligonucleotide primed in situ labeling (PRINS): applications in pathology. Histochem.<br />

Cell <strong>Bio</strong>l. 108, 439– 446.<br />

13. Herrington, C. S. and O’Leary, J. J. (1998) PCR In Situ Hybridization: A Practical<br />

Approach, Oxford University Press, Oxford, U.K.<br />

14. Speel, E. J. M., Hopman, A. H. N., and Komminoth, P. (1999) Amplification methods to<br />

increase the sensitivity of in situ hybridization. Play card(s). J. Histochem. Cytochem.<br />

47, 281–288.<br />

15. Darby, I. A. (2000) In Situ Hybridization Protocols. Methods in Molecular <strong>Bio</strong>logy, Vol.<br />

123, Humana Press, Totowa, NJ.<br />

16. Coullin, P., Roy, L., Pellestor, F., Candelier, J.-J., Bed”hom, B., Guillier-Gencik, Z., et al.<br />

(2002) PRINS, the other in situ DNA labeling method that is useful in cellular biology.<br />

Am. J. Med. Genet., in press.<br />

17. Hindkjœr, J., Bolund, L., and Kolvraa, S. (2001) Primed in situ labeling, in Methods in Cell<br />

<strong>Bio</strong>logy, Vol. 63–64, Academic Press, San Diego, pp. 55–68.<br />

18. Speel, E. J. M., Lawson, D., Ramaekers, F. C. S., Gosden, J. R., and Hopman, A. H. N.<br />

(1996) Rapid brightfield detection of oligonucleotide primed in situ (PRINS)-labeled DNA<br />

in chromosome preparations and frozen tissue sections. <strong>Bio</strong>Techniques 20, 226–234.<br />

19. Ramael, M., Van Steelandt, H., Stuyven, G., Van Steenkiste, M., and Degroote, J. (1999)<br />

Detection of human papilloma virus (HPV) genomes by the primed in situ labelling<br />

technique. Pathol. Res. Pract. 195, 801–807.<br />

20. Gosden, J. and Lawson, D. (1994) Rapid chromosome identification by oligonucleotideprimed<br />

in situ DNA synthesis (PRINS). Hum. Mol. Genet. 3, 931–936.<br />

21. Koch, J., Hindkjœr, J., Kolvraa, S., and Bolund, L. (1995) Construction of a panel of<br />

chromosome-specific oligonucleotide probes (PRINS-primers) useful for the identification<br />

of individual human chromosomes in situ. Cytogenet. Cell. Genet. 71, 142–147.<br />

22. Pellestor, F., Girardet, A., Lefort, G., Andréo, B., and Charlieu, J.-P. (1995) Selection of<br />

chromosome specific primers and their use in simple and double PRINS technique for rapid<br />

in situ identification of human chromosomes. Cytogenet. Cell. Genet. 70, 138–142.<br />

23. Roguel-Gaillard, C., Hayes, H., Coullin, P., Chardon, P., and Vaiman, M. (1997) Swine<br />

centromeric DNA repeats revealed by primed in situ (PRINS) labeling. Cytogenet. Cell<br />

Genet. 79, 79–84.<br />

24. Gosden, J., Hanratty, D., Starling, J., Fantes, J., Mitchell, A., and Porteous, D. (1991) Oligonucleotide-primed<br />

in situ DNA synthesis (PRINS): A method for chromosome mapping,<br />

banding, and investigation of sequence organization. Cytogenet. Cell. Genet. 57, 100–104.


422 Speel, Ramaekers, and Hopman<br />

25. Krejci, K. and Koch, J. (1999) An in situ study of variant telomeric repeats in human<br />

chromosomes. Genomics 58, 202–206.<br />

26. Go, Y., Rakotoarisoa, G., Kawamoto, Y., Randrianjafy, A., Koyama, N., and Hirai, H. (2000)<br />

PRINS analysis of telomeric sequence in seven lemurs. Chrom. Res. 8, 57–65.<br />

27. Cinti, C., Santi, S., and Maraldi, N. M. (1993) Localization of a single copy gene by PRINS<br />

technique. Nucleic Acids Res. 21, 5799–5800.<br />

28. Kadandale, J. S., Tunca, Y., and Tharapel, A. T. (2000) Chromosomal localization of single<br />

copy genes SRY and SOX3 by primed in situ labeling (PRINS). Microbial. Comparative<br />

Genomics 5, 71–74.<br />

29. Harrer, T., Schwinger, E., and Mennicke, K. (2001) A new technique for cyclic in situ<br />

amplification and a case report about amplification of a single copy gene sequence in human<br />

metaphase chromosomes through PCR-PRINS. Hum. Mutat. 17, 131–140.<br />

30. Pellestor, F. (2002) PRINS and cycling PRINS on chromosomes and sperm cells, in<br />

Chromosome Analysis Protocols, 2nd ed, Methods in Molecular <strong>Bio</strong>logy (Speel, E. J. M.,<br />

and Hopman, A. H. N., eds.), Humana Press, Totowa, NJ, in press.<br />

31. Volpi, E. V. and Baldini, A. (1993) MultiPRINS: A method for multicolor primed in situ<br />

labeling. Chrom. Res. 1, 257–260.<br />

32. Hindkjœr, J., Koch, J., Terkelsen, C., Brandt, C. A., Kolvraa, S., and Bolund, L. (1994) Fast,<br />

sensitive multicolor detection of nucleic acids in situ by primed in situ labeling (PRINS).<br />

Cytogenet. Cell. Genet. 66, 152–154.<br />

33. Speel, E. J. M., Lawson, D., Hopman, A. H. N., and Gosden, J. (1995) Multi-PRINS:<br />

multiple sequential oligonucleotide primed in situ DNA synthesis reactions label specific<br />

chromosomes and produce bands. Hum. Genet. 95, 29–33.<br />

34. Schröck, E., Du Manoir, S., Veldman, T., Schoell, B., Wienberg, J., Ferguson-Smith, M. A.,<br />

et al. (1996) Multicolor spectral karyotyping of human chromosomes. Science 273,<br />

494– 497.<br />

35. Speicher, M. R., Ballard, S. G., and Ward, D. C. (1996) Karyotyping human chromosomes<br />

by combinatorial multi-fluor FISH. Nat. Genet. 12, 368–375.<br />

36. Schriml, L. M., Padilla-Nash, H. M., Coleman, A., Moen, P., Nash, W. G., Menninger, J.,<br />

et al. (1999) Tyramide signal amplification (TSA)-FISH applied to mapping PCR-labeled<br />

probes less than 1 kb in size. <strong>Bio</strong>Techniques 27, 608–613.<br />

37. Terkelsen, C., Koch, J., Kolvraa, S., Hindkjœr, J., Pedersen, S., and Bolund, L. (1993)<br />

Repeated primed in situ labeling: formation and labeling of specific DNA sequences in<br />

chromosomes and nuclei. Cytogenet. Cell Genet. 63, 235–237.<br />

38. Troyer, D. L., Goad, D. W., Xie, H., Rohrer, G. A., Alexander, L. J., and Beattle, C.<br />

W. (1994) Use of direct in situ single-copy (DISC) PCR to physically map five porcine<br />

microsatellites. Cytogenet. Cell Genet. 67, 199–204.<br />

39. Tharapel, S. A. and Kadandale, J. S. (2002) Primed in situ labeling (PRINS) for evaluation<br />

of gene deletions in cancer. Am. J. Med. Genet., in press.<br />

40. Paskins, L., Brownie, J., and Bull, J. (1999) In situ polymerase chain reaction and cycling primed<br />

in situ amplification: improvements and adaptations. Histochem. Cell <strong>Bio</strong>l. 111, 411– 416.<br />

41. Jonveaux, J. (1991) PCR amplification of specific DNA sequences from routinely fixed<br />

chromosomal spreads. Nucleic Acids Res. 19, 1946.<br />

42. Höfler, H. (1993) In situ polymerase chain reaction: toy or tool? Histochemistry 99,<br />

103–104.<br />

43. Teo, I. A. and Shaunak, S. (1995) Polymerase chain reaction in situ: An appraisal of an<br />

emerging technique. Histochem. J. 27, 647–659.<br />

44. Hopman, A. H. N. and Ramaekers, F. C. S. (1998) Processing and staining of cell and tissue<br />

material for interphase cytogenetics, in Current Protocols in Cytometry (Robinson, J. P.,<br />

ed.), <strong>John</strong> Wiley & Sons, New York, pp. 8.5.1–8.5.22.


PRINS and In Situ PCR 423<br />

45. Sällström, J. F., Zehbe, I., Alemi, M., and Wilander, E. (1993) Pitfalls of in situ polymerase<br />

chain reaction (PCR) using direct incorporation of labelled nucleotides. Anticancer Res.<br />

13, 1153.<br />

46. Speel, E. J. M., Hopman, A. H. N., and Komminoth, P. (2000) Signal amplification for DNA<br />

and mRNA in situ hybridization, in In Situ Hybridization Protocols. Methods in Molecular<br />

<strong>Bio</strong>logy (Darby, I. A., ed.), Humana Press, Totowa, NJ, pp. 195–216.<br />

47. Sperry, A., Jin, L., and Lloyd, R. V. (1996) Microwave treatment enhances detection of<br />

RNA and DNA by in situ hybridization. Diagn. Mol. Pathol. 5, 291–296.<br />

48. Pilling, A., Endersby-Wood, H., and Jones, S. (1997) Increased sensitivity of in situ<br />

hybridization: implications for in situ amplification. Diagn. Mol. Pathol. 6, 174.<br />

49. Nuovo, G. J. (2001) Co-labeling using in situ PCR: A review. J. Histochem. Cytochem.<br />

49, 1329–1339.<br />

50. Komminoth, P., Long, A. A., Ray, R., and Wolfe, H. J. (1992) In situ polymerase chain<br />

reaction detection of viral DNA, single copy genes and gene rearrangements in cell<br />

suspensions and cytospins. Diagn. Mol. Pathol. 1, 85–97.<br />

51. Wiedorn, K. H., Kühl, H., Galle, J., Caselitz, J., and Vollmer, E. (1999) Comparison of in<br />

situ hybridization, direct and indirect in situ PCR as well as tyramide signal amplification<br />

for the detection of HPV. Histochem. Cell <strong>Bio</strong>l. 111, 89–95.<br />

52. Patterson, B., Till, M., Otto, P., Goolsby, C., Furtado, M., McBride, L., and Wolinsky, S.<br />

(1993) Detection of HIV-1 DNA and messenger RNA in individual cells by PCR-driven in<br />

situ hybridization and flow cytometry. Science 260, 976–979.<br />

53. Speel, E. J. M., Richter, J., Moch, H., Egenter, C., Saremaslani, P., Rütimann, K., et al.<br />

(1999) Genetic differences in endocrine pancreatic tumor subtypes detected by comparative<br />

genomic hybridization. Am. J. Pathol. 155, 1787–1794.<br />

54. Zehbe, I., Hacker, G. W., Su, H., Hauser-Kronberger, C., Hainfeld, J. F., and Tubbs, R.<br />

(1997) Sensitive in situ hybridization with catalyzed reporter deposition, streptavidinnanogold,<br />

and silver acetate autometallography. Am. J. Pathol. 150, 1553–1561.<br />

55. Lizardi, P. M., Huang, X., Zhu, Z., Bray-Ward, P., Thomas, D. C., and Ward, D. C.<br />

(1998) Mutation detection and single-molecule counting using isothermal rolling-circle<br />

amplification. Nat. Genet. 19, 225–232.<br />

56. Höfler, H., Pütz, B., Mueller, J., Neubert, W., Sutter, G., and Gais, P. (1995) In situ<br />

amplification of measles virus RNA by the self-sustained sequence replication reaction.<br />

Lab. Invest. 73, 577–585.<br />

57. Tecott, L. H., Barchas, J. D., and Eberwine, J. H. (1988) In situ transcription: specific<br />

synthesis of complementary DNA in fixed tissue sections. Science 240, 1661–1664.<br />

58. Mogensen, J., Kolvraa, S., Hindkjœr, J., Petersen, S., Koch, J., Nygård, M., et al. (1991)<br />

Nonradioactive, sequence-specific detection of RNA in situ by primed in situ labeling<br />

(PRINS). Exp. Cell Res. 196, 92–98.<br />

59. Player, A. N., Shen, L.-P., Kenny, D., Antao, V. P., and Kolberg, J. A. (2001) Single-copy<br />

gene detection using branched DNA (bDNA) in situ hybridization. J. Histochem. Cytochem.<br />

49, 603–611.<br />

60. Mangham, D. C. and Isaacson, P. G. (1999) A novel immunohistochemical detection<br />

system using mirror image complementary antibodies (MICA). Histopathology 35,<br />

129–133.<br />

61. Pastore, J. N., Clampett, C., Miller, J., Porter, K., and Miller, D. (1995) A rapid immunoenzyme<br />

double labeling technique using enzyme polymer system (EPOS) reagents.<br />

J. Histotechnol. 18, 35–40.<br />

62. Kämmerer, U., Kapp, M., Gassel, A. M., Richter, T., Tank, C., Dietl, J., and Ruck, P.<br />

(2001) A new rapid immunohistochemical staining technique using the EnVision antibody<br />

complex. J. Histochem. Cytochem. 49, 623–630.


424 Speel, Ramaekers, and Hopman<br />

63. Van Gijlswijk, R. P. M., Van Gijlswijk-Janssen, D. J., Raap, A. K., Daha, M. R., and<br />

Tanke, H. J. (1996) Enzyme-labeled antibody-avidin conjugates: new flexible and sensitive<br />

immunochemical reagents. J. Immunol. Methods 189, 117–127.<br />

64. Chen, B.-X., Szabolcs, M. J., Matsushima, A. Y., and Erlanger, B. F. (1996) A strategy for<br />

immunohistochemical signal enhancement by end-product amplification. J. Histochem.<br />

Cytochem. 44, 819–824.


Cycling Primed In Situ Amplification 425<br />

60<br />

Cycling Primed In Situ Amplification<br />

<strong>John</strong> H. Bull and Lynn Paskins<br />

1. Introduction<br />

Primed in situ amplification (PRINS) is a technique for the visualization of specific<br />

sequences, usually repeat sequences, in fixed cell nuclei. When viewed on the microscope,<br />

the resulting signals can be seen as spots within nuclei, providing a means to<br />

visualize telomeres, centromeric regions, Alu repeats, or other sequences.<br />

At its simplest, the PRINS reaction is a primer extension conducted under a sealed<br />

coverslip on a microscope slide. A mix of oligonucleotide primer, dNTPs (including<br />

a labeled dNTP), and Taq DNA polymerase is applied to a preparation of fixed cell<br />

nuclei on the slide, which is placed on a heating block and subjected to a round of<br />

denaturation, annealing, and extension. During this process, the nuclei are held in place<br />

while the DNA strands are made available for oligonucleotide annealing and extension<br />

by the polymerase. Unincorporated nucleotides are washed off, and the incorporated<br />

labeled nucleotide is detected typically by fluorescence (1,2).<br />

PRINS has many applications in common with the widely used fluorescence in situ<br />

hybridization (FISH) technique (3). However, PRINS has significant advantages in the<br />

speed of the reaction, avoidance of toxic chemicals, lack of dependence on a carefully<br />

controlled stringency wash step, and in the use of easily synthesized oligonucleotides<br />

rather than expensive probes. For example, specific primers can be used for most<br />

human chromosomes or pairs of chromosomes (4), giving discrete subnuclear spots<br />

that allow chromosome enumeration. The target sequence is typically satellite repeat<br />

(e.g., α-satellite, a family of 171-bp repeat units present at the centromeres of human<br />

chromosomes) and the strength of the signal is a function of the number of repeat<br />

units at the target site.<br />

Cycling PRINS represents a modification where the primer extension is repeated<br />

several times (see Fig. 1), with the result that multiple labeled DNA strands are<br />

synthesized with a concomitant increase in signal (5). This has obvious advantages in<br />

sensitivity, but in practice there are constraints to the signal build-up in cycling PRINS,<br />

which center around the fixation method used to prepare the nuclei or cells. PRINS<br />

depends on unimpeded access of reagents to the nuclear DNA. As the newly primed<br />

strand is synthesized, it is held in place by base pairing to the nuclear template DNA.<br />

When the reaction is cycled, there may be nothing to stop the newly synthesized strand<br />

diffusing away from the site of synthesis. In our laboratory, this problem appears<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

425


426 Bull and Paskins<br />

Fig. 1. Results of cycling PRINS using α-satellite primers for specific chromosomes (see<br />

Subheading 3.1.), DIG-dUTP incorporation and FITC detection. (A) cultured lymphoblast<br />

cells settled onto a slide labeled for chromosome X, (B) peripheral blood smear also labeled for<br />

chromosome X, and (C) smear of bone marrow aspirate labeled for chromosome 7. See Note 4<br />

for details of cell preparation.


Cycling Primed In Situ Amplification 427<br />

most severe when the cytoplasm has been stripped away, as in typical methanol/acetic<br />

acid fixed metaphase spread preparations used in cytogenetic analysis (6). The naked<br />

nucleus provides no barrier to diffusion, and although PRINS is very efficient, cycling<br />

PRINS confers no advantage as a steady state of signal is reached after the first cycle.<br />

Many elaborate schemes can be imagined to overcome this, involving crosslinking<br />

agents, looped or circular structures, and photoactivatable anchors. The method<br />

described here simply uses ethanol fixation of whole cells. When ethanol is used,<br />

signal retention in cycling PRINS is markedly improved (7). Background labeling<br />

of the nuclear DNA is increased with this fixative, but this is easily counteracted by<br />

microwave boiling of the slides before the PRINS reaction is set up. Using an extended<br />

initial denaturation step also helps reduce background.<br />

The method described here is applicable to cell preparations, such as blood smears,<br />

cytospin, or gravity-settled cell preparations. Once dry, the slides can be fixed and stored<br />

for several months before cycling PRINS analysis. Although it is presently limited<br />

to whole cell preparations and seems incompatible with the analysis of metaphase<br />

spreads, it is simple, robust, and effective.<br />

2. Materials<br />

2.1. Slide Preparation<br />

1. Glass slides and coverslips: high-grade, dust-free slides are a sound investment and<br />

require no further preparation (e.g., Menzel Glaser Super Premium, Fisher Scientific,<br />

Loughborough, UK). Coated slides can be useful if there are doubts around loss of valuable<br />

cells (see Note 1). Coverslips can be 22 × 22 mm or bigger.<br />

2. Cell preparation (see Subheading 3.2.).<br />

3. Ethanol (100%).<br />

4. Slide staining jar, for example, a Coplin jar.<br />

5. 10 mM Tris-HCl, 5 mM EDTA, pH 7.0.<br />

6. Anti-bumping granules (Merck).<br />

7. 800-W Microwave oven (e.g., Matsui M162TC).<br />

8. Graded ethanol solutions: 100%, 90%, and 70% v/v with water.<br />

2.2. Cycling PRINS<br />

1. AmpliTaq Gold DNA polymerase (PE <strong>Bio</strong>systems).<br />

2. 10× PCR buffer: 500 mM KCl, 100 mM Tris-HCl, pH 8.3, 200 mM MgCl2 (see Note 2).<br />

3. dNTPs (Amersham Pharmacia <strong>Bio</strong>tech Inc., Uppsala, Sweden). Dilute to make a 50× stock<br />

solution of 10 mM each dATP, dCTP, dGTP; and 1.8 mM dTTP.<br />

4. Digoxygenin-11-dUTP (DIG-dUTP; 10 nmol/µL, Roche Molecular <strong>Bio</strong>chemicals, Lewes,<br />

UK) (see Note 3).<br />

5. Oligonucleotide primer (50 µM, see Subheading 3.1.).<br />

6. Double distilled water.<br />

7. High-grade clean coverslips and rubber solution (see Note 4) or Amplicovers and<br />

Ampliclips (PE <strong>Bio</strong>systems).<br />

8. Flat-bed or slide-adapted thermal cycling (“PCR”) machine (e.g., Hybaid, Ashford, UK<br />

or PE <strong>Bio</strong>sytems).<br />

9. Stop buffer: 500mM NaCl, 50 mM EDTA, pH 7.0.<br />

10. Water bath at 65°C.


428 Bull and Paskins<br />

2.3. Detection<br />

1. Water bath at 45°C, and staining jar (e.g., a Coplin jar).<br />

2. Incubator set at 37°C and sandwich box.<br />

3. 20× SSC: 3.0 M NaCl, 0.30 M tri-sodium citrate, pH 7.3.<br />

4. Wash buffer: 4× SSC (diluted from 20× stock), 0.05% Triton×-100.<br />

5. Dried skimmed milk powder.<br />

6. Blocking buffer: Wash buffer with the addition of 5% skimmed milk powder (which can<br />

be stored for a few days at 4°C).<br />

7. Anti DIG-FITC (Roche Molecular <strong>Bio</strong>chemicals). Store at –20°C in 5-µL aliquots.<br />

8. Antifade mountant (e.g., Vectashield, Vector Laboratories, Burlingame CA).<br />

9. Counterstain: propidium iodide solution (20 µg/mL, Sigma, Dorset, UK) (see Note 5).<br />

10. Fluorescence microscope with appropriate filters for FITC and PI (e.g., Zeiss Axioskop ® ,<br />

Carl Zeiss, Thornwood, NY).<br />

3. Methods<br />

3.1. Primer Design<br />

These protocols were tested using α-satellite-specific primers for DYZ1 (D599:<br />

TGGGCTGGAATGGAAAGGAATCGAAAC), DXZ1 (E563: ATAATTTCCCATA-<br />

ACTAAACACA), D17Z1 (E571: AATTTCAGCTGACTAAACA), D7Z1 (E528:<br />

AGCGATTTGAGGACAATTGC), and D3Z1 (E570: TCTGCAAGTGGATATTTAAA)<br />

(1). It has been suggested that one or more of these are used to set up the technique<br />

when using human cells (see Note 5).<br />

3.2. Cell Preparation<br />

The type of cells you use will of course depend on your experimental system. Aim to<br />

prepare cells in a monolayer, preferably slightly separated from each other, with several<br />

hundred cells in an area of 20 to 30 mm 2 for ease of location at the microscopy stage<br />

(see Note 5). In practice, 10 to 20 cells may be all that is needed for good results.<br />

Human peripheral blood leukocytes provide a robust positive control and, satisfyingly,<br />

are easily available in plentiful and cheap supply (see Note 6). The majority red<br />

cells do not interfere with the PRINS reaction.<br />

1. Obtain a few drops of blood from your finger using a suitable puncture device (e.g..<br />

Haemolance, HaeMedic AB, Munka Ljungby, Sweden).<br />

2. Spot approx 5 µL whole blood onto one end of a slide. Use a second microscope slide to<br />

spread the blood. This is done by holding the end of the second slide at an angle of roughly<br />

45 degrees onto the blood spot, allowing it to spread under the edge, then pushing with<br />

some force to smear the cells. When done properly, a monolayer or “feather” of cells a few<br />

mm wide will form at the end of the smear (see Note 5). With practice, it is easily possible<br />

to make 2 or 3 small smears per slide, each using 1 to 2 µL of blood. This is useful if space<br />

on the flat-block PCR machine is limiting.<br />

3. Allow to air dry.<br />

3.3. Fixation<br />

1. Fix by immersion in ethanol (see Note 7) for 5 min at room temperature (1 min is sufficient<br />

for blood smears).<br />

2. Remove slides, drain, and air dry. Slides can be used immediately or stored at 4°C for<br />

several months.


Cycling Primed In Situ Amplification 429<br />

3.4. Microwave Pretreatment<br />

1. Place about 10 anti-bumping granules in a 50-mL Coplin jar and add the microscope slides.<br />

Slides can be placed back to back so that a jar contains 10 slides.<br />

2. Fill the jar with 10 mM Tris-HCl, 5 mM EDTA (40–50 mL).<br />

3. Place the jar in the center of the microwave oven, switch on full power until boiling,<br />

and boil for a further 50 s.<br />

4. Quickly transfer the slides to a staining jar containing 70% ethanol at room temperature.<br />

Leave for 1 min.<br />

5. Pass through 90% and 100% ethanol for 1 min each. Slides can either be air dried for<br />

immediate use or stored at 4°C in 100% ethanol for several months if desired.<br />

3.5. Cycling PRINS<br />

Reaction volumes will vary with your system (see Note 8). As a rough guide,<br />

22- × 22-mm coverslips require about 15 µL, and 50- × 22-mm coverslips or Amplicovers<br />

require about 40 µL.<br />

1. Prepare 100-µL reaction mix as follows: 2 µL of dNTP mix, 0.4 µL of DIG-dUTP, 10 µL of<br />

10× PCR buffer, 2 µL (10 units) of AmpliTaq Gold DNA polymerase, 2 µL oligonucleotide<br />

primer, and 83.6 µL water.<br />

2. Place 15 to 40 µL on the slide area, according to the cell preparation area and size of your<br />

coverslip, or 40 µL for Amplicovers.<br />

3. For coverslips, seal with rubber solution and allow this to dry (a fan or laminar flow cabinet<br />

speed this up). For Amplicovers, follow manufacturers’ instructions.<br />

4. Transfer the slides to a flat block thermal cycler. A suitable program for the primers<br />

described is: 18 min at 95°C (to activate the AmpliTaq Gold DNA polymerase: reduce<br />

to 5 min for standard Taq), then 94°C for 1 min, 60°C for 1 min, and 72°C for 1 min<br />

for 15 cycles (see Note 8).<br />

5. Peel off rubber glues, or take off Amplicovers, and immerse slides in a Coplin jar containing<br />

stop buffer at 65°C. Coverslips may come off with the glue, if not they will fall off readily<br />

in the stop bath. After 1 min, transfer slides to a jar containing wash buffer (we have also<br />

used 2× SSC, 45°C for 5 min to stop the reaction).<br />

3.6. Detection<br />

Do not allow slides to dry during this process.<br />

1. Prepare blocking buffer, and pipet 40 µL on to a clean cover slip. Shake the slide free<br />

of excess wash buffer, then pick up the cover slip with slide. Leave the slide at room<br />

temperature for 5 min.<br />

2. Dilute the anti-DIG FITC 1/100 in blocking buffer (e.g., add 500 µL of blocking buffer<br />

to a 5-µL aliquot. This is usually a suitable dilution, although batch-to-batch variation<br />

may be encountered).<br />

3. Remove the coverslip from the slide and drain the excess fluid by shaking or briefly<br />

blotting one edge against an absorbent paper towel. Pipet 40 µL of diluted anti-DIG FITC<br />

on to the coverslip and replace the slide. Incubate in a moist chamber (e.g., a sandwich<br />

box lined with damp filter paper) at 37°C for 30 min.<br />

4. Wash the slides 3× for 2 min in prewarmed wash buffer (42°C) in a Coplin jar.<br />

5. Prepare mountant: add 3.75 µL of propidium iodide (see Note 9) stock to 100 µL of<br />

Vectashield.


430 Bull and Paskins<br />

6. Shake the slide free of excess liquid and mount as follows: pipet 40 µL of mountant on<br />

to a clean coverslip and pick up the coverslip with the slide. Place between two layers of<br />

tissue and press to spread the mountant and expel the excess. Seal with rubber solution<br />

(for long term storage) and allow to dry.<br />

7. Slides can be viewed at once or stored for up to a few months in the dark at 4°C without<br />

significant loss of signal. Typical results are shown in Fig. 1.<br />

4. Notes<br />

1. Coated slides are often used in histological procedures where harsh treatments can cause<br />

loss of sample from the slide surface. They can be bought from Fisher or PE <strong>Bio</strong>systems<br />

or prepared by using poly L-Lysine or aminopropyltriethoxysilane (8). The Amplicover ®<br />

and Ampliclip ® system, together with slides, supplied by PE <strong>Bio</strong>systems (Foster City,<br />

CA) provides an alternative option for sealing the reaction solution onto the slide surface,<br />

but also requires use of their In Situ PCR 1000 thermal cycler and tool for assembling<br />

the slide chamber. Although relatively expensive, this is the best system for more than<br />

five cycles of PRINS.<br />

2. The relatively high concentration of MgCl 2 is to counteract loss of solution phase Mg<br />

thought to occur during cycling.<br />

3. Alternatively, biotin-labeled dUTP can be used at the same concentration, in which case<br />

a suitable detection reagent is avidin-FITC (Vector Labs). This can be substituted for<br />

anti-DIG FITC, at a 1500 dilution.<br />

4. Most commercially available brands for cycle puncture repair perform adequately. Rema<br />

Tip Top (Munich, Germany) is recommended.<br />

5. Hematological smears work well, as do cytocentrifuge preparations of lymphocytes or<br />

neutrophils or cultured cells. Cells prepared by these methods are flattened and have a<br />

distinct morphological appearance, which is an advantage in interpreting and recording<br />

results because subnuclear spots tend to be in the same focal plane. We have used standard<br />

hematological preparations, such as bone marrow or peripheral blood smears, Giemsastained<br />

slides, and archival slides stored at room temperature, for over a year without<br />

problems. For cultured cells, or leukocyte preparations made from fresh blood (e.g.,<br />

Lymphoprep, Sigma), cytocentrifuge (“cytospin”) preparations are ideal (e.g., Shandon or<br />

Hiraeus cytocentrifuges and equipment). Alternatively, cells can be allowed to settle on<br />

coated slides for 10 to 20 min then allowed to dry after draining off liquid. These methods<br />

are easy to perform, but molecular biologists unfamiliar with cytological preparations<br />

might benefit from a tutorial in a hospital hematology or histology laboratory. Crystallized<br />

solids from support medium or isotonic solutions are not usually problematic because they<br />

are removed in the fixation and microwaving procedure. For extra phenotypic <strong>info</strong>rmation,<br />

immunocytochemical staining can be done, and signals can be visualized alongside PRINS<br />

signal (7,9).<br />

6. Hygiene precautions (i.e., hand washing) should be taken. The implications of sampling<br />

one’s own blood might merit consideration: An inadvertent diagnosis of aneuploidy may<br />

cause distress! Chromosomes 3, 7, and 17 should be safe in this regard.<br />

7. Ethanol fixation works by precipitating proteins irreversibly from solution. Other precipitating<br />

fixatives we have successfully used include methanol and acetone.<br />

8. Reactions can be performed under coverslips or using the PE Applied <strong>Bio</strong>systems system<br />

(Amplicovers and Ampliclips). If using coverslips, some slides may start to dry out after<br />

around 10 cycles, but this will depend on the size of your coverslip and the type of rubber<br />

solution used to seal it to the slide. We find that 5 to 7 cycles is a good compromise. Using<br />

the PE system, we have cycled up to 70 times.


Cycling Primed In Situ Amplification 431<br />

9. Propidium iodide binds specifically to DNA and emits a red fluorescence distinct from<br />

FITC. If desired, DAPI (4′,6-Diamidino-2-phenylindole 2 HCl; Sigma) can also be used<br />

at the same concentration. This emits a strong blue fluorescence, which sometimes aids in<br />

locating the cells on the slide. As an alternative detection system, Texas Red-conjugated<br />

reagents are available for detection of DIG or biotin. This can be combined with the DAPI<br />

counterstain, leaving the green channel free for detection of a third label, for example,<br />

via immunostaining (7,9).<br />

References<br />

1. Gosden, J. and Lawson, D. (1994) Rapid chromosome identification by oligonucleotideprimed<br />

in situ synthesis (PRINS). Hum. Mol. Genet. 3, 931–936.<br />

2. Hindkjaer, J., Koch, J., Terkelsen, C., Brandt, C. A., Kolvraa, S., and Bolund, S. (1994)<br />

Fast, sensitive multicolour detection of nucleic acids in situ by primed in situ labelling<br />

(PRINS). Cytogenet. Cell Genet. 66, 152–154.<br />

3. Trask, B. J. (1991) Fluorescence in situ hybridization: Applications in cytogenetics and<br />

gene mapping. Trends Genet. 7, 149–54.<br />

4. Koch, J., Hindkjaer, J., Kolvraa, S., and Bolund, L. (1995) Construction of a panel of<br />

chromosome-specific oligonucleotide probes (PRINS-primers) useful for the identification<br />

of individual human chromosomes in situ. Cytogenet. Cell Genet. 71, 142–147.<br />

5. Gosden, J. and Hanratty, D. (1993) PCR in situ: A rapid alternative to in situ hybridisation<br />

for mapping short, low-copy number sequences without isotopes. <strong>Bio</strong>Techniques 15,<br />

78–83.<br />

6. Spowart, G. (1994) Mitotic metaphase chromosome preparation from peripheral blood for<br />

high resolution, in Methods in Molecular <strong>Bio</strong>logy, Vol. 29: Chromosome Analysis Protocols<br />

(Gosden, J. R., ed.), Humana, Totowa, NJ, pp. 1–10.<br />

7. Paskins, L., Brownie, J., and Bull, J. (1999) In situ polymerase chain reaction and cycling<br />

primed in situ amplification: improvements and adaptations. Histochem. Cell <strong>Bio</strong>l. 111,<br />

411– 416.<br />

8. Jackson, P. and Blyth, D. (1993) Immunolabelling techniques for light microscopy,<br />

in Immuncytochemistry, a Practical Approach (Beesley, J. E., ed.), IRL, Oxford, UK.<br />

pp. 15– 42.<br />

9. Speel, J. M., Lawson, D., Ramaekers, C. S., Gosden, J. R., and Hopman, A. H. N. (1997)<br />

Combined immunocytochemistry and PRINS DNA synthesis for simultaneous detection<br />

of phenotypic and genomic parameters in cells, in Methods in Molecular <strong>Bio</strong>logy, Vol. 71:<br />

PRINS and in Situ PCR Protocols (Gosden, J.R., ed.), Humana, Totowa, NJ, pp. 53–60.


432 Bull and Paskins


Direct and Indirect In Situ PCR 433<br />

61<br />

Direct and Indirect In Situ PCR<br />

Klaus Hermann Wiedorn and Torsten Goldmann<br />

1. Introduction<br />

In recent years, the development of in situ technologies has made good progress.<br />

In situ hybridization (ISH) has become an important tool and has enabled the pathologist<br />

to demonstrate infectious pathogens or mRNAs in tissue sections or cytospins without<br />

destruction of morphology, thus enabling the assignment of signals to individual cells<br />

or cell compartments (1–9).<br />

Although ISH has contributed substantially to the diagnosis and understanding of<br />

neoplastic and infectious diseases, the detection of low copy DNA and RNA sequences<br />

by conventional ISH remained difficult in the past because of the relatively low<br />

sensitivity of ISH, irrespective of whether the investigation was performed using<br />

radioactive or nonradioactive hybridization probes (8,9a,10–15). Nonradioactive<br />

probes, especially biotin and digoxigenin (DIG), which are less hazardous to work<br />

with, can be much more quickly developed, allow a much higher spatial resolution,<br />

and have been shown to be at least as sensitive as radioactive probes (10–14). Several<br />

different detection systems (14,16–20) have been used to enhance signal intensities.<br />

Nevertheless, conventional ISH usually will enable detection of high to medium copy<br />

number nucleic acids only.<br />

However, target amplification by in situ PCR (IS-PCR) or reverse transcription<br />

in situ PCR (RT-IS-PCR) have been shown to even allow the detection of low copy<br />

number DNAs or RNAs (1,4–6,8,15,21–23). There exist two different approaches to<br />

IS-PCR: direct and indirect IS-PCR (Fig. 1).<br />

Indirect IS-PCR requires an additional ISH step and is more cumbersome but will<br />

usually yield reliable results. Direct IS-PCR is often hampered by nonspecific products,<br />

especially when performed on paraffin-embedded tissue sections. These false positives<br />

in direct IS-PCR are predominantly primer-independent artifacts resulting from DNA<br />

repair and endogenous priming (5,21,22,24,25). These pathways are also operative in<br />

indirect IS-PCR but will not produce false positives because no labeled nucleotides are<br />

incorporated during the amplification step. Primer-dependent artifacts like mispriming<br />

can be controlled by hot start maneuvers, although this is in general not quite as easy<br />

as in solution-phase PCRs. In addition, diffusion artifacts may be involved in the<br />

generation of false-positive cells in paraffin-embedded tissues undergoing direct as well<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

433


434 Wiedorn and Goldmann<br />

Fig. 1. IS-PCR can be performed with indirect and direct approach. With direct IS-PCR<br />

during amplification, unlabeled and labeled nucleotides are used in the reaction mix. Thus,<br />

labeled nucleotides are incorporated into the PCR products. Therefore, subsequent to the<br />

amplification, direct detection of the labeled PCR products can be performed. In contrast with<br />

indirect IS-PCR, only unlabeled nucleotides are used with the reaction mix. Therefore, the<br />

amplicons are unlabeled, and for detection of the PCR products subsequent to the amplification,<br />

an ISH with a labeled internal probe (a probe that does not contain the primer sites) has to be<br />

performed. Thus, the detection cannot be performed until after ISH is finished.<br />

as indirect IS-PCR, although they are predominantly associated with cell suspensions<br />

(4,22). Optimized fixation and permeabilization as well as reduction of PCR cycle<br />

numbers to less than 30 cycles are likely to minimize but not totally exclude this<br />

phenomenon. With RT-IS-PCR, in contrast to solution-phase PCR, an additional<br />

problem will arise. Depending on the primers chosen, false-positive nuclear signals<br />

may arise because usually the primers will anneal to cDNA as well as to genomic<br />

DNA in the nucleus (which is eliminated during RNA extraction and additional DNAse<br />

digestion for removing residual DNA in approaches using solution phase RT-PCR) and<br />

will therefore not only generate PCR-products in the cytoplasm but also in the nucleus.<br />

To circumvent this problem, cDNA-specific primers, which will only anneal to the<br />

cDNA, may be designed for RT-IS-PCR (15). In addition, a DNAse pretreatment may<br />

be performed before RT-IS-PCR to destroy the genomic DNA. However, the general<br />

potential of this technique is controversially discussed (5,15,26).<br />

Nonetheless, although direct IS-PCR will contain more risks for false-positive<br />

results, both methods are described in this chapter especially because direct IS-PCR<br />

offers a more convenient way and will yield reliable results when performed on<br />

nonparaffin-embedded samples like cytospins.<br />

2. Materials<br />

1. In situ Thermal cycling machine (Hybaid AGS, Heidelberg, Germany; Shandon, Frankfurt,<br />

Germany) (see Note 1).<br />

2. Wash Module (Hybaid AGS, Heidelberg, Germany; Shandon, Frankfurt, Germany).<br />

3. SuperFrostPlus slides (Menzel Gläser, Braunschweig, Germany) (see Note 2).


Direct and Indirect In Situ PCR 435<br />

4. Cover Slips (Omnilab, Germany).<br />

5. Microtome (Leica, Germany).<br />

6. Heating Oven for 220°C (WTB Binder, Germany).<br />

7. Water bath (e.g., GFL 1083, GFL, Germany).<br />

8. Pattex Supermatic 200plus (Henkel, Düsseldorf, Germany).<br />

9. Forceps.<br />

10. Laboratory gas burner.<br />

11. DEPC (Sigma, Deisenhofen, Germany, Cat-No. D5758).<br />

12. 5 NNaOH (Merck, Germany).<br />

13. 1 NHCl (Merck, Germany).<br />

14. NaCl (Merck, Germany).<br />

15. NaH 2 PO 4 H 2 O (Merck, Germany).<br />

16. Na 2 HPO 4 2H 2 O (Merck, Germany).<br />

17. Tris-HCl (Sigma, Deisenhofen, Germany).<br />

18. EDTA (Sigma, Deisenhofen, Germany).<br />

19. 10× phosphate-buffered saline (PBS): 1.3 M NaCl, 5 mM NaH 2 PO 4 , 95 mM Na 2 HPO 4 .<br />

Adjust to pH 7.2 with orthophosphoric acid.<br />

20. 5% buffered formalin pH 7.2 (Frontell, Germany).<br />

21. 4% buffered paraformaldehyde, pH 7.2. Dissolve 40 g of paraformaldehyde (Merck ultra<br />

pure 4005, Germany) in 1× PBS. Adjust to pH 7.2 with NaOH and fill up to 1000 mL with<br />

1× PBS. Filter through Wathman No 1.<br />

22. 100% ETOH (Merck, Germany).<br />

23. Xylene (Merck, Germany).<br />

24. Taq Polymerase 5 U/µL (Roche Diagnostic GmbH, Mannheim, Germany, Cat No.<br />

1146173).<br />

25. 0.1 M Tris-HCl, 0.1 M NaCl, pH 7.5.<br />

26. 1 M MgCl 2 (Sigma, Deisenhofen, Germany).<br />

27. CaCl 2 (C3306, Sigma, Deisenhofen, Germany).<br />

28. 20× SSC: 3 M NaCl and 0.3 M Na 3 Citrate. Adjust to pH 7.0 with HCl.<br />

29. Bovine serum albumin (BSA) 20 mg/ mL (Roche Diagnostic GmbH, Mannheim, Germany,<br />

Cat-No. 711454).<br />

30. 10% SDS (Sigma, Deisenhofen, Germany, Cat-No. L 4522).<br />

31. 100% deionized formamide (Sigma, Deisenhofen, Germany, Cat-No. F9037).<br />

32. Dextran sulfate (Sigma, Deisenhofen, Germany, Cat-No. D8906): 50% dextran sulfate.<br />

Dissolve 50 mg in 100 mL of DEPC-treated distilled water. Heat to 60°C and shake to<br />

speed up the dissolving.<br />

33. DIG Wash and Block Buffer Set (Roche Diagnostic GmbH, Mannheim, Germany, Cat-No.<br />

1585762) containing 10× washing buffer, 10× blocking solution (=10% blocking solution),<br />

10× detection buffer, 10× maleic acid buffer.<br />

34. NBT 100 mg/mL (Nitroblue tetrazolium chloride; Roche Diagnostic GmbH, Mannheim,<br />

Germany Cat-No. 1383213).<br />

35. BCIP 50 mg/mL (5-Brom-4-chlor-3-indoyl-phosphate) (Roche Diagnostic GmbH, Mannheim,<br />

Germany, Cat-No. 138221).<br />

36. Mounting media (as supplied; Permount SP15-500, Fisher Scientific, Wiesbaden,<br />

Germany).<br />

37. Counterstain (Nuclear Fast Red and Eosin, Merck, Germany).<br />

38. Anti-Dig or Anti-biotin antibody conjugates (Roche Diagnostic GmbH, Mannheim,<br />

Germany Cat-No. 1426303, 1426311, 1093274, 1207733).<br />

39. Hybridization probes (Eurogentec, Belgium or MWG <strong>Bio</strong>tech, Ebersberg, Germany).<br />

40. DNAse (Roche Diagnostic GmbH, Mannheim, Germany, Cat-No. 776785).


436 Wiedorn and Goldmann<br />

41. Proteinase K solution: Proteinase K (Roche Diagnostic GmbH, Mannheim, Germany;<br />

250 µg/mL), 100 mM Tris-HCl and 50 mM EDTA.<br />

42. 10× Target Retrieval (Dako, Hamburg, Germany, Cat-No. S1700).<br />

43. DNAse solution: 40 mM Tris-HCl, pH 7.4, 6 mM MgCl 2 , 2 mM CaCl 2 , and 1 U/µL RNAse<br />

free DNAse (Roche Diagnostic GmbH, Mannheim, Germany).<br />

44. IL6 PCR mix: 10 mM Tris-HCl, pH 9.2, 50 mM KCl, 3 mM MgCl 2 , 0.2 mM dNTPs<br />

(Roche Diagnostic GmbH, Mannheim, Germany); 0.1% BSA (Roche Diagnostic GmbH,<br />

Mannheim, Germany); 10 µM DIG-11-dUTP (Roche Diagnostic GmbH, Mannheim,<br />

Germany) (see Note 3); 5 U/sample Taq-Polymerase (Roche Diagnostic GmbH, Mannheim,<br />

Germany); and 0.4 µM of sense and antisense primer.<br />

45. Sequence of IL6 primers (see Note 4): sense primer: 5′ CTTCTCCACAAGCGCCTTC-3′;<br />

antisense primer: 3′ CTAAGTTACTCCTCTGAACGG-5′.<br />

46. A typical hybridization mix is composed of (exact compositions depends on the probe in<br />

use): 2× up to 5× SSC; up to 50% formamide; 5% up to 10% dextran sulfate (see Note 5);<br />

0.1% SDS; 0.1% BSA; 1% up to 2% blocking solution; 250 µg/ mL fish sperm DNA<br />

(Roche Diagnostic GmbH, Mannheim, Germany); 1–10 ng/µL hybridization probe.<br />

Additionally for RT-IS-PCR:<br />

47. DNAse/RNAse solution: 40 mM Tris-HCl, pH 7.4, 6 mM MgCl 2 , 2 mM CaCl 2 , 1 U/µL<br />

RNAse free DNAse (Roche Diagnostic GmbH, Mannheim, Germany), and500 µg/mL<br />

RNAse (Roche Diagnostic GmbH, Mannheim, Germany).<br />

48. Reverse transcription buffer (RT-buffer): 50 mM Tris-HCl, pH 8.3, 75 mM KCl, 3 mM<br />

MgCl 2 , 0.5 mM dNTPs (Roche Diagnostic GmbH, Mannheim, Germany), 10 mM DTT<br />

(GibcoBRL, Karlsruhe, Germany), 1 U/µL RNAsin (Promega, Mannheim, Germany),<br />

2 U/µL M-MLV-RT (GibcoBRL, Karlsruhe, Germany) (see Note 6), 0.1 µg/µL Random<br />

Hexamers (Roche Diagnostic GmbH, Mannheim, Germany).<br />

3. Methods<br />

If not otherwise mentioned, all incubations are performed at ambient temperature<br />

(20°C). Incubations on the thermocycler are performed using the humidity chamber.<br />

Samples are processed without coverslips and sealing if not otherwise mentioned.<br />

3.1. Prevention of DNAse and RNAse Contamination<br />

1. Incubate all slides, coverslips and microtome blades for 12 h at 220°C to inactivate DNAse<br />

and RNAse (see Notes 7 and 8).<br />

2. If RNA detection is required, include 0.1% DEPC in all solutions and incubate overnight.<br />

3. Autoclave solutions for 20 min at 1.2 bar (15 psi) at 121°C.<br />

3.2. Fixation<br />

Incubate tissues for 24 h in 5% buffered formalin (for detection of DNA) or 4%<br />

buffered paraformaldehyde (for detection of mRNA) (see Note 9). With cell suspensions<br />

and cytospins a 30 min incubation is usually sufficient.<br />

3.3. Paraffin Embedding<br />

The distilled water used to dilute ETOH should be autoclaved and prepared with<br />

DEPC-treated water to ensure the absence of DNAse and RNAse.<br />

1. Incubate tissue for 30–60 min in 70% ETOH.<br />

2. Incubate tissue for 30–60 min in 80% ETOH.


Direct and Indirect In Situ PCR 437<br />

3. Incubate tissue for 30–60 min in 90% ETOH.<br />

4. Incubate tissue for 30–60 min in 96% ETOH.<br />

5. Incubate tissue for 30–60 min in 100% ETOH.<br />

6. Incubate tissue for 30–60 min in 100% ETOH.<br />

7. Incubate tissue for 60 min in xylene100% ETOH (11).<br />

8. Incubate tissue 2 × 60 min in xylene.<br />

9. Incubate tissue 3 × 60 min in paraffin.<br />

10. Allow to cool and harden the paraffin for several hours.<br />

3.4. Preparation of Samples<br />

3.4.1. Tissue Sections<br />

1. Use Teflon coated SuperFrostPlus slides (see Note 2).<br />

2. Cut section of 2- to 5-µm thickness (see Note 10).<br />

3. Change the microtome blade frequently and clean thoroughly with xylene after each block<br />

to prevent crosscontamination between samples.<br />

4. Incubate slides overnight at 50°C.<br />

3.4.2. Cytospins<br />

1. Centrifuge cells on slides.<br />

2. Incubate slides for at least 30 min at 50°C.<br />

3.5. Deparaffinization<br />

1. Incubate slides for 10 min at 70°C.<br />

2. Incubate slides for 10 min in fresh xylene (see Note 11).<br />

3. Incubate slides for 2 min in fresh xylene.<br />

4. Incubate 2 × 1 min in 100% ETOH.<br />

5. Incubate 1 min in 90% ETOH.<br />

6. Incubate 1 min in 70% ETOH.<br />

7. Incubate 1 min in 50% ETOH.<br />

8. Incubate 1 min in distilled water.<br />

9. Dry Slides for up to 10 min at 37°C (on the thermocycler).<br />

3.6. Permeabilization<br />

The following permeabilization protocols are for optimized fixation conditions and<br />

may vary with fixation conditions and tissues (see Notes 9 and 12).<br />

3.6.1. By Protease<br />

1. Incubate 15 min at 37°C (on the thermocycler) with 100 µL of Proteinase K solution.<br />

2. Incubate 2 × 5 min in 0.1 M Tris-HCl, 0.1 M NaCl, pH 7.5.<br />

3.6.2. By Target Retrieval<br />

1. Put slides in a Coplin jar filled with 1 × Target Retrieval (Dako). Place Coplin jar in a<br />

waterbath and incubate 35 min at 95°C.<br />

2. Place Coplin jar outside the water bath and allow to cool for 20 min.<br />

3. Incubate 2 × 5 min in 0.1 M Tris-HCl, 0.1 M NaCl, pH 7.5.<br />

3.7. DNAse Treatment<br />

1. Incubate in 100 µL of DNAse solution at 37°C for 12 h on the thermocycler. Cover slides<br />

with coverslips.


438 Wiedorn and Goldmann<br />

2. Incubate 4 min at 95°C.<br />

3. Incubate 2 × 5 min in 0.1 M Tris-HCl, 0.1 M NaCl, pH 7.5, in a Coplin jar.<br />

3.8. DNAse/RNAse Treatment<br />

If destruction of RNAse is necessary too (for example to generate negative controls),<br />

a combined DNAse/RNAse treatment is recommended. Proceed as mentioned under<br />

Subheading 3.7., but use the DNAse/RNAse solution.<br />

3.9. Quenching<br />

If Peroxidase/AEC or -/DAB systems are used during the detection then quenching<br />

is required to inactivate endogenous peroxidase.<br />

1. Incubate 30 to 45 min in 100 µL of 0.6% hydrogen peroxide (see Note 13).<br />

2. Incubate 1 min in 50% ETOH.<br />

3. Incubate 1 min in 70% ETOH.<br />

4. Incubate 1 min in 90% ETOH.<br />

5. Dry Slides for 5 min at 37°C (on the thermocycler).<br />

3.10. Reverse Transcription<br />

For all RT and IS-PCR, the Omnislide and MISHA thermocyclers have to be set to<br />

simulated slide control and calibration factor 100.<br />

If you are investigating DNA targets, then proceed to Subheading 3.11.<br />

1. Incubate for 60 min at 39°C in 15 to 20 µL of RT-buffer per sample on the thermocycler.<br />

Cover sample with coverslips and seal with PattexSupermatic200Plus (see Note 14).<br />

2. Stop reaction by incubation at 92°C for 10 min.<br />

3. Remove coverslips (see Note 15) and discard reaction mixture.<br />

3.11. IS-PCR<br />

Use aerosol resistant pipet tips for preparing the PCR.<br />

1. Prepare PCR mix without Taq-Polymerase in Eppendorf tubes.<br />

2. Heat PCR mix to 80°C.<br />

3. Heat slides to 80°C on the thermocycler.<br />

4. Add Taq-Polymerase to PCR mix and vortex briefly.<br />

5. Pipet 16 µL of the PCR mix onto the specimen.<br />

6. Immediately cover with a coverslip using a forceps and seal with PattexSupermatic<br />

200Plus.<br />

7. Heat forceps in a laboratory gas burner before proceeding to the next sample to avoid<br />

cross contamination.<br />

8. Repeat steps 5–7 for each specimen.<br />

9. Start the PCR program.<br />

10. Remove coverslips (see Note 15) and discard reaction mixture.<br />

The PCR protocol for IL6 is composed of one cycle at 92°C for 5 min, 32 cycles each<br />

with 1 min denaturation step at 92°C, 1 min annealing step at 56°C, 2 min extension<br />

step at 72°C. The program is finished by an additional extension for 7 min at 72°C.<br />

3.12. ISH<br />

This step is only necessary if indirect IS-PCR is performed. Otherwise proceed to<br />

Subheading 3.13. All reactions are performed on the thermocycler.


Direct and Indirect In Situ PCR 439<br />

1. Prepare fresh hybridization mix without fish sperm DNA.<br />

2. Denature fish sperm DNA for 10 min at 100°C in a water bath.<br />

3. Cool fish sperm DNA on ice for 2 min.<br />

4. Add fish sperm DNA to the hybridization mix.<br />

5. Pipet 16 µL of the hybridization mix onto each sample.<br />

6. Immediately cover with a coverslip using a forceps and seal with PattexSupermatic200Plus<br />

(see Note 14).<br />

7. Heat forceps in a laboratory gas burner before proceeding to the next sample to avoid<br />

cross contamination.<br />

8. Incubate 10 min at 95°C.<br />

9. Incubate 1 to 16 h between 30 to 65°C (see Note 16).<br />

10. Remove coverslips (see Note 15) and discard reaction mixture.<br />

3.13. Washing and Detection<br />

1. Incubate 2 × 5 min in 0.1× SSC at 20°C (see Note 17).<br />

2. Incubate 10 min in 0.1× SSC at 45°C in the Wash Module (see Note 17).<br />

3. Incubate 1 min in 1× washing buffer (DIG Wash and Block buffer Set, Roche Diagnostic<br />

GmbH, Mannheim, Germany).<br />

4. Incubate 60 min in 1× blocking solution at 37°C.<br />

5. Add 100 µL of 1250 diluted Anti-DIG-AP-conjugate per sample and incubate 60 min<br />

at 37°C on the thermocycler.<br />

6. Incubate 2 × 5 min in 1× washing buffer at 20°C.<br />

7. Incubate 2 × 5 min in 1× detection buffer at 20°C.<br />

8. Incubate in 100 µL per sample freshly made substrate solution (up to 12 h) in the<br />

dark. Substrate solution is composed of (per mL): 4.5 µL NBT, 3.5 µL BCIP, 992 µL<br />

of detection buffer (all reagents Roche Diagnostic GmbH, Mannheim, Germany) (see<br />

Note 18).<br />

9. Counterstain 10 min with nuclear fast red or eosin at 20°C (intensity of counterstain may<br />

be adjusted by reducing or prolonging incubation period).<br />

10. Mount with Permount SP15-500 (Fisher Scientific).<br />

3.14. Controls<br />

Several controls have to be performed to ensure the specificity of the results.<br />

1. Omission of RT (should result in cells showing no signal).<br />

2. Omission of Taq-Polymerase (should result in cells showing no signal).<br />

3. Omission of primers (should result in cells showing no signal).<br />

4. RT IS-PCR of a housekeeping gene, such as GAPDH, or actin to ensure that the reverse<br />

transcription and PCR was successful because GAPDH should be demonstrable in each<br />

cell. Furthermore, this should indicate if the permeabilization was successful.<br />

5. PCR of Alu-repeats to ensure that PCR and permeabilization was successful.<br />

6. Omission of the Anti-Dig-POD or Anti-DIG-AP to detect endogenous enzyme activity.<br />

7. Tissues that are definitively negative for the sequence under investigation. If such tissues<br />

are not available, the control samples may be treated either with DNAse, RNAse or both.<br />

8. Tissues that are definitively positive for the sequence under investigation.<br />

Furthermore, at least in the phase of establishing a new PCR protocol for new<br />

primers, all IS-PCRs should be controlled by simultaneous solution-phase PCRs on<br />

serial sections of the same samples.


440 Wiedorn and Goldmann<br />

3.15. Primer Design<br />

Primers should be designed to be cDNA specific (if RNA targets are under investigation),<br />

that is, to span an intron (15) to disable amplification of genomic sequences<br />

during RT-IS-PCR. Furthermore, primers should be chosen to give small amplification<br />

products because the efficiency of IS-PCR will be reduced with longer products.<br />

3.16. Closing Remarks<br />

PCR has become an important diagnostic as well as research tool in molecular biology,<br />

clinical chemistry, and pathology. With the invention of IS-PCR, the amplification<br />

power of solution-phase PCR with no limitations in the amount of template was hoped<br />

to be transferred to the in situ techniques. Unfortunately, this has become reality only<br />

partly because IS-PCR is often hampered by poor reproducibility, specificity, and<br />

reliability (8,9) and by the cumbersome protocol. For semiquantitative in situ studies of<br />

gene expression in combination with image analysis, RT-IS-PCR seems to be of little<br />

value because of the tremendously varying amplification efficiency of IS-PCR (15,27).<br />

For these applications, ISH with subsequent signal amplification by biotinyl tyramide<br />

proved to be the method of choice. This approach has been shown to be an excellent<br />

alternative for IS-PCR. With respect to most applications, generally signal amplification<br />

procedures are more suitable than target amplification by direct or indirect IS-PCR<br />

and exhibit a sensitivity similar to that of IS-PCR (2,8,9,25,28). Adequate choice<br />

of hybridization probes provided signal amplification allows even the detection of<br />

single-copy virus sequences (28).<br />

4. Notes<br />

1. We have made good experience with the Omnislide in situ Thermocycler (Hybaid AGS<br />

Germany) and the MISHA (Shandon Germany) because the slides are fitted horizontally<br />

onto the blocks in the humidity chamber thus requiring sealing only during the PCR<br />

whereas those apparatus where slides have to be fitted vertically onto the blocks require<br />

sealing during each step of the procedure resulting in a much more cumbersome protocol.<br />

For achieving reliable results, it is of utmost importance to use thermocyclers specially<br />

designed for in situ PCR procedures. For detailed instructions regarding the setup of the<br />

thermocycler and protocols, see instruction manual and (29).<br />

2. Although expensive (E 0.50 per slide) we recommend the use of Teflon coated SuperFrost-<br />

Plus slides (Menzel-Gläser, Braunschweig, Germany). These behaved well with respect to<br />

adhesion of tissue even after prolonged cycling protocols and because of the hydrophobic<br />

Teflon coating around the well do not require the cumbersome use of hydrophobic pens to<br />

outline the reaction area. Hydrophobic pens usually have to be used several times during<br />

a PCR protocol to guarantee a closed border around the tissue sample. However, repeated<br />

application of the hydrophobic pens requires drying of the slide and the tissue, which can<br />

produce strong background staining.<br />

Slides with at least two wells should be used so that controls can be run on the same<br />

slide simultaneously. In this case, Teflon coating will effectively prevent contamination<br />

between the two samples. We used slides with wells of 17-mm diameter each. The volumes<br />

indicated during the protocol proofed well in covering samples of this size but have to be<br />

adjusted for wells with different diameters.<br />

3. DIG-dUTP is omitted if indirect IS-PCR is performed. Instead of DIG-dUTP, biotinlabeled<br />

nucleotides can be used too.<br />

4. Unlike in solution-phase, PCR primers for IS-PCR should be designed to give amplicons


Direct and Indirect In Situ PCR 441<br />

of 200 to 800 bp because with larger amplicons the amplification efficiency will be reduced<br />

dramatically and will often result in false negative samples.<br />

5. Because dextran sulfate may produce background staining, especially if higher concentrations<br />

are used, reduce the amount of dextran sulfate to diminish the background staining.<br />

For ISH in contrast to filter hybridization, usually 10% dextran sulfate is the upper limit.<br />

6. The reverse transcription can also be performed with other transcriptases, such as<br />

Superscript (GibcoBRL, Karlsruhe, Germany). In this case, the protocol has to be adjusted<br />

accordingly.<br />

7. For SuperFrostPlus slides, this is only a precaution as when stored (dust free) and<br />

handled under appropriate conditions, we have encountered neither DNAse nor RNAse<br />

contamination.<br />

8. Coverslips may be siliconized prior to the incubation at 220°C if adherence of tissue<br />

sample to the coverslips is observed during the PCR procedure.<br />

9. To obtain reproducible results, it is essential that fixation is always performed at the same<br />

temperature and for the same time. Otherwise, the permeabilization conditions have to<br />

be adapted and optimized for each PCR. A difference of 5°C during fixation may give<br />

false negative results because of changes in cell permeability. Unfortunately, tissue in<br />

molecular pathology is usually not fixed under standardized conditions. Therefore, several<br />

permeabilization conditions have to be tested for each sample.<br />

To improve fixation and therefore DNA and RNA preservation, tissues should be cut<br />

to the minimum size prior to fixation.<br />

For mRNA detection paraformaldehyde will yield a better preservation of RNA than<br />

formalin. Perform fixation at 4°C where possible. The extraction of RNA from formalinor<br />

paraformaldehyde-fixed tissues for doing a solution-phase PCR to verify the in situ<br />

results especially when establishing a new protocol is usually very ineffective. For these<br />

applications, we recommend the use of the Hope fixative (patent pending, Dr. Olert, Institut für<br />

Kinderpathologie, Universität Mainz, Germany available at DCS, Hamburg, Germany) which<br />

additionally yields a much better preservation of RNA compared to other fixatives.<br />

10. Thinner sections show better adhesion to slides during IS-PCR, although this reduces the<br />

number of target sequences, which may result in some negative cells.<br />

11. Use fresh reagents for each incubation. Because xylenes saturate rapidly with paraffin use<br />

200 mL of fresh xylene for each batch of 15 to 20 slides.<br />

12. Although fixed under the same conditions, different tissues may require adapted permeabilization<br />

parameters. When optimizing permeabilization parameters with respect to<br />

morphology, it is usually better to prolong Proteinase K incubation than to increase the<br />

concentration of Proteinase K (15). If diffusion artifacts are a major problem the reduction<br />

of the concentration of Proteinase K with simultaneous prolongation of incubation turned<br />

out to be advantageous (15). If you encounter background staining this may be reduced<br />

by the use of target retrieval instead of Proteinase K (15). Proteinase K (250 µg/ mL) is<br />

optimal for portio and condylomata biopsies fixed for 24 h in buffered formalin whereas<br />

for leukocytes 10 to 30 µg/mL Proteinase K and for bronchial epithelial cells 30 to<br />

50 µg/mL Proteinase K performed well. Leukocytes and bronchial epithelial cells were<br />

fixed with 4% buffered paraformaldehyde for 30 min.<br />

13. The concentration of endogenous peroxidase varies significantly between tissues. Leukocytes<br />

exhibiting high endogenous peroxidase levels usually require longer incubation<br />

(45 min) than other tissues. If you encounter signals in your negative controls try to<br />

diminish those signals either by prolongation of the quenching reaction or by increasing<br />

the concentration of H 2 O 2 up to 3%.<br />

14. Evaporation control is of the utmost importance for reliable and reproducible results as<br />

unspecific signals because of the evaporation of the reaction mixture may result from insuf-


442 Wiedorn and Goldmann<br />

ficient sealing. From the many methods for sealing, the use of Pattex Supermatic200Plus<br />

offers a convenient and reliable way to prevent leakage (15,29,30). It is easier to handle<br />

than nail polish with the same good sealing capacity. Furthermore, it seems to represent a<br />

biocompatible adhesive as it does not interfere with enzyme activity during PCR.<br />

15. Coverslips can easily be removed by an initial scalpel cut along the sealed edge, which<br />

is to be lifted first. To facilitate the removal of the coverslips, the slides can be stored at<br />

4°C for 1 to 2 min to completely harden the gluten if it is still viscous after the incubation<br />

at 92°C (8,15,29).<br />

16. The lower the temperature and the shorter the incubation period, the better will be<br />

morphology which, nevertheless, is negatively affected by the IS-PCR procedure.<br />

17. Temperatures, concentration of SSC, and incubation period have to be adjusted for the<br />

probe in use.<br />

18. Antibody concentrations have to be adapted for different probes and according to the<br />

efficiency of the PCR. One can choose from a wide variety of chromogens. Furthermore, a<br />

peroxidase-based system can be used using AEC+ or DAB+ (Dako) as chromogens, which<br />

usually will result in more localized signals (8,28) than those achieved with NBT/BCIP.<br />

Detection solutions and antibody concentrations have to be adjusted.<br />

Acknowledgments<br />

The authors are indebted to Prof. Dr. Dr. E. Vollmer for making it possible to participate<br />

in the development of this technology and H. Kühl for excellent technical assistance.<br />

References<br />

1. Chen, R. H. and Fuggle, S. V. (1993) In situ cDNA polymerase chain reaction. A novel<br />

technique for detecting mRNA expression. Am. J. Pathol. 143, 1527–1534.<br />

2. Goldmann, T., Becher, B., Wiedorn, K. H., Pirow, R., Deutschbein, M. E., Vollmer, E.,<br />

et al. (1999) Epipodite and fat cells as sites of hemoglobin synthesis in the branchiopod<br />

crustacean Daphnia magna. Histochem. Cell <strong>Bio</strong>l. 112, 335–339.<br />

3. Heniford, B. W., Shum-Siu, A., Leonberger, M., and Hendler, F. J. (1993) Variation in<br />

cellular, E. G.F receptor mRNA expression demonstrated by in situ reverse transcriptase<br />

polymerase chain reaction. Nucleic Acids Res. 21, 3159–3166.<br />

4. Long, A. A., Komminoth, P., and Wolfe, H. J. (1992) Detection of HI.V provirus by in situ<br />

polymerase chain reaction. N. Engl. J. Med. 327, 1529.<br />

5. Nuovo, G. J. (1994) PCR in Situ Hybridization Protocols and Applications, Second Ed.,<br />

Raven Press, New York.<br />

6. Patel, V. G., Shum-Siu, A., Heniford, B. W., Wieman, T. J., and Hendler, F. J. (1994)<br />

Detection of epidermal growth factor receptor mRNA in tissue sections from biopsy<br />

specimens using in situ polymerase chain reaction. Am. J. Pathol. 144, 7–14.<br />

7. Snijders, P. J. F., van den Brule, A. J. C., Schrijnemakers, H. F. J., Snow, G., Meijer, C.<br />

J. L. M., and Walboomers, J. M. M. (1990) The use of general primers in the polymerase<br />

chain reaction permits detection of a broad spectrum of human papillomavirus genotypes.<br />

J. Gen. Virol. 71, 173–181.<br />

8. Wiedorn, K. H., Kühl, H., Galle, J., Caselitz, J., and Vollmer, E. (1999) Comparison of insitu<br />

hybridization, direct and indirect in-situ PCR as well as tyramide signal amplification<br />

for the detection of HPV. Histochem. Cell <strong>Bio</strong>l. 111, 89–95.<br />

9. Wiedorn, K. H., Goldmann, T., Kühl, H., Galle, J., and Vollmer, E. (1999) GenPoint :<br />

A new biotinyl tyramide based signal amplification system for significantly increasing<br />

sensitivity of in situ hybridization (ISH). Elec. J. Pathol. Histol. 994–999.


Direct and Indirect In Situ PCR 443<br />

9a. Furuta, Y., Shinohara, T., Sano, K., Meguro, M., and Nagashima, K. (1990) In situ<br />

hybridization with digoxigenin-labelled DNA probes for detection of viral genomes.<br />

J. Clin. Pathol. 43, 806–809.<br />

10. Allan, G. M., Todd, D., Smyth, J. A., Mackie, D. P., Burns, J., and McNulty, M. S. (1989) In<br />

situ hybridization: an optimised detection protocol for a biotinylated DNA probe renders it<br />

more sensitive than a comparable 35 S-labeled probe. J. Virol. Methods 24, 181–190.<br />

11. Burns, J., Graham, A., Frak, C., Fleming, F., Evans, M., and McGee, J. (1987) Detection of<br />

low copy human papilloma virus DNA and mRNA in routine paraffin sections of cervix by<br />

non-isotopic in situ hybridization. J. Clin. Pathol. 40, 858–864.<br />

12. Furuta, Y., Shinohara, T., Sano, K., Meguro, M., and Nagashima, K. (1990) In situ<br />

hybridization with digoxigenin-labelled DNA probes for detection of viral genomes.<br />

J. Clin. Pathol. 43, 806–809.<br />

13. Nuovo, G. J. and Richart, R. M. (1989) A comparison of biotin and 35 S-based in situ<br />

hybridization methodologies for detection of human papilloma-virus DNA. Lab. Invest.<br />

61, 471–476.<br />

14. Syrjänen, S., Partanen, P., Mäntyjärvi, R., and Syrjanen K. (1988) Sensitivity of in situ<br />

hybridization techniques using biotin- and 35S-labeled human papillomavirus (HPV) DNA<br />

probes. J. Virol. Methods 19, 225–238.<br />

15. Wiedorn, K. H., Lang, D., Köhl, H., Galle, J., and Vollmer, E. (1999) Sensitive in situ<br />

detection of modulated expression levels of interleukin 6 (IL6) and interleukin 8 (IL8) in<br />

bronchial epithelium and leucocytes using cDNA specific primers and direct RT in-Situ-<br />

PCR. Elec. J. Pathol. Histol. 993–996.<br />

16. Herrington, C. S., Burns, J., Graham, A. K., Evans, M., and McGee, J. O. (1989) Interphase<br />

cytogenetics using biotin and digoxigenin labeled probes I: relative sensitivity of both<br />

reporter molecules for detection of HPV 16 in CaSki cells. J. Clin. Pathol. 42, 592–600.<br />

17. Holm, R., Karlsen, F., and Nesland, J. M. (1992) In situ hybridization with nonisotopic<br />

probes using different detection systems. Mod. Pathol. 5, 315–319.<br />

18. Lassus, J., Niemi, K.-M., Marjamaki, A., Syrjanen, S., Kartamaa, M., Lehmus, A., et al.<br />

(1992) Comparison of four in situ hybridization methods, based on digoxigenin- and<br />

biotin-labelled probes, in detecting HPV DNA in male condylomata acuminata. Int.<br />

J. STD AIDS 3, 196–203.<br />

19. Mifflin, T. E., Bowdwn, J., Lovell, M. A., Burns, D. E., Hayden, F. G., Graschel, D. H. M,<br />

et al. (1987) Comparison of radioactive and biotinylated DNA probes for the detection of<br />

cytomegalovirus. Clin. <strong>Bio</strong>chem. 20, 231–235.<br />

20. Niedobitek, G., Finn, T., Herbst, H., and Stein, H. (1989) In situ hybridization using<br />

biotinylated probes. An evaluation of different detection systems. Pathol. Res. Pract.<br />

184, 343–348.<br />

21. Komminoth, P. and Long, A. A. (1993) In situ polymerase chain reaction. An overview<br />

of methods, applications and limitations of a new molecular technique. Virchows Archiv.<br />

64, 67–73.<br />

22. Komminoth, P. and Long, A. A. (1995) In situ polymerase chain reaction—methodology,<br />

applications and nonspecific pathways, in PCR Applications Manual, Boehringer Mannheim,<br />

Mannheim, pp. 97–106.<br />

23. Long, A. A., Komminoth, P., and Wolfe, H. J. (1993) Comparison of indirect and direct in<br />

situ polymerase chain reaction in cell preparations and tissue sections. Detection of viral<br />

DNA gene rearrangements and chromosomal translocations. Histochemistry 99, 51–162.<br />

24. Söllström, J. F., Zehbe, I., Alemi, M., and Wilander, E. (1993) Pitfalls of in situ polymerase<br />

chain reaction (PCR) using direct incorporation of labelled nucleotides. Anticancer Res.<br />

13, 1153–1154.


444 Wiedorn and Goldmann<br />

25. Komminoth, P. and Werner, M. (1997) Target and signal amplification: approaches to<br />

increase the sensitivity of in situ hybridization. Histochem. Cell <strong>Bio</strong>l. 108, 325–333.<br />

26. Komminoth, P., Adams, V., Long, A. A., Roth, J., Saremaslani, P, Flury, R., et al. (1994)<br />

Evaluation of methods for hepatitis C virus (HCV) detection in liver biopsies: Comparison<br />

of histology, immunohistochemistry, in situ hybridization, reverse (RT) PCR and in situ RT<br />

PCR. Pathol. Res. Pract. 190, 1017–1025.<br />

27. Wiedorn, K. H., Schaaf, B., Kühl, H., Spuck, S., Galle, J., Dalhoff, K., et al. (1999)<br />

Granulocyte Colony Stimulating Factor (G-CSF) induced modulation of cytokine expression:<br />

A semiquantitative in situ analysis using target amplification by reverse transcriptase<br />

in situ PCR (RT IS PCR) as well as signal amplification by in situ hybridization (ISH)<br />

with catalyzed reporter deposition (CARD). Verhandlungen der Deutschen Gesellschaft<br />

für Zytologie 21, 63–65.<br />

28. Wiedorn, K. H., Goldmann, T., Kühl, H., Rohrer, Ch., and Vollmer, E. (2000) Single-copy<br />

virus detection by GenPoint in situ hybridisation on the Eppendorf Mastercycler. <strong>Bio</strong>Tech<br />

Int. 12, 12–13.<br />

29. Wiedorn, K. H. (1996) In situ Hybridisierungshandbuch zum MISHA in situ Thermocycler,<br />

in Misha Multipler In-Situ-Hybridisierungsautomat für die In-Situ-Hybridisierung und<br />

Polymerase-Kettenreaktion (PCR), Shandon, Frankfurt.<br />

30. Stapleton, M. J., Levin, M. C., and Jacobson, S. (1994) DNA Amplification within cells on<br />

slides: Advances in evaporation and temperature control. Cell Vision 1, 177–181.


RT In Situ PCR 445<br />

62<br />

Reverse Transcriptase In Situ PCR<br />

New Methods in Cellular Interrogation<br />

Mark Gilchrist and A. Dean Befus<br />

1. Introduction<br />

The advent of the reverse transcriptase polymerase chain reaction (RT-PCR)<br />

technique represents a quantum leap in sensitivity over preceding methods of detecting<br />

mRNA transcripts, such as Northern blotting. With the arrival of such sensitive<br />

techniques, it has become possible to amplify RNA transcripts from very small amounts<br />

of template nucleic acid, thus opening new avenues of research that were previously<br />

off limits because of difficulties in obtaining adequate quantities and quality of RNA<br />

(1,2). However, RT-PCR suffers from the same limitations as its predecessor because<br />

the isolation of RNA necessitates the destruction of the cells/tissue involved, thus<br />

preventing the identification of the specific cell source of the mRNA (3). Conversely,<br />

in situ hybridization allows the specific localization of mRNA to the cells of origin,<br />

but the methodology is much less sensitive than RT-PCR (3). A methodology that<br />

combines the best attributes of in situ hybridization (specific cellular localization) and<br />

RT-PCR (high sensitivity) would be desirable. RT in situ PCR provides these attributes,<br />

allowing for the location and detection of low copy RNA species, amplified within<br />

individual intact cells (4).<br />

The method (Fig. 1) involves fixation of cells in suspension, followed by controlled<br />

digestion of the crosslinked cellular proteins with a proteolytic enzyme. This allows<br />

the entry of the RT and PCR reagents into the cell. Next, genomic DNA is removed<br />

by a DNase digestion step, thus ensuring that only mRNA is amplified. Subsequent<br />

steps of RT and PCR are then undertaken, using a “labeled” reporter nucleotide, which<br />

results in its direct incorporation into the PCR product. This is followed by a detection<br />

step that visualizes the amplified mRNA, either by chromogenic or radioactive means<br />

(Fig. 1). The specificity of the reaction can be then established by several methods<br />

(see Note 1).<br />

Methods of RT in situ PCR, although sharing fundamental steps, have varied greatly<br />

between laboratories (5–7). On the basis of our experience, RT in situ PCR seems to be<br />

best suited for the detection of mRNA in single-cell suspensions, in which fixation and<br />

pretreatments can be optimally controlled (8). The method outlined below is similar to<br />

that developed by Nuovo et al. 4,5) and has been adopted and built upon through our<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

445


446 Gilchrist and Befus<br />

Fig. 1. Schematic representation of the steps involved in the general protocol outlined in<br />

the text. After fixation of the cells in suspension, the crosslinks that formed during fixation are<br />

reduced by limited protease digestion. To remove the possibility of amplifying DNA during<br />

the PCR, a DNase step is then performed to remove contaminating genomic DNA. A reverse<br />

transcription reaction is then performed to convert the mRNA to cDNA. PCR amplification with<br />

gene specific primers is then completed. The PCR product is then detected by chromogenic,<br />

fluorescent or radioactive means. Many commonly used RT in situ PCR procedures follow<br />

similar guidelines, differing in specific details, such as fixative used, protease used, digestion<br />

time, and method of detection.<br />

experiences. The inclusion of appropriate controls in every run helps insure accurate<br />

results. Controls to indicate that genomic DNA has been removed as a source of<br />

false-positive signals (negative control), or failure of PCR amplification because<br />

of inadequate protease digestion or flaws in the RT-PCR protocol resulting in falsenegative<br />

results (positive control) must be tested on every slide. Therefore, a slide<br />

schematic of the necessary controls and test spots is included to aid in the elimination<br />

of spurious results (Fig. 2).<br />

2. Materials<br />

1. Phosphate-buffered saline (PBS): 130 mM NaCl, 10 mM sodium phosphate, pH 7.4. Store<br />

at room temperature.<br />

2. 10% neutral buffered formalin (BDH). Store at room temperature.<br />

3. Heparinase I (for heparin containing mast cell populations; Sigma).<br />

4. Heparinase buffer: 5 mM Tris-HCl, pH 7.5, 1 mM CaCl 2 , 7.5 U RNasin.<br />

5. Pepsin (5000 U/mL), made fresh in 0.01 N HCl for immediate use (Boehringer<br />

Mannheim).<br />

6. RNase-free DNase I (Boehringer Mannheim).<br />

7. DNase solution: DNase I (10 U/µL) in 0.1 M sodium acetate, pH 5.0, 5 mM MgSO 4 .


RT In Situ PCR 447<br />

Fig. 2. Schematic of the appropriate controls, how they are achieved, and the expected results<br />

that indicate a successful reaction.<br />

8. Primers (see Subheading 4.6.). The following primer pair yields a 295-bp amplicon for rat<br />

TNF mRNA: 5-TACTGAACTTCGGGGTGATCGGTCC-3 and 5-CAGCCTTGTCCCTT<br />

GAAGAGAACC-3.<br />

9. M-MLV RT enzyme (Gibco/BRL).<br />

10. Reverse transcription buffer: 75 mM KCl, 3 mM MgCl 2 , 50 mM Tris-HCl, 0.1 M DTT<br />

(pH 8.3), 1 µM antisense primer or 25 µg/mL oligo-(dT) 12–18 primer, and 2500 U/mL<br />

M-MLV RT enzyme.<br />

11. Taq DNA polymerase (Gibco/BRL).<br />

12. PCR buffer: 50 mM KCl, 10 mM Tris-HCl (pH 8.3), 4.5 mM MgCl 2 , 80 µM mixed dNTP,<br />

16 µM digoxigenin-11-dUTP, 1.2 µM each primer, and 120 U/mL Taq polymerase.<br />

13. Digoxigenin-11-dUTP (Boehringer Mannheim).<br />

14. Anti-Digoxigenin antibody, alkaline phosphatase conjugated (Boehringer Mannheim).<br />

15. Wash buffer 1: 0.1 M TRIS, 0.15 M NaCl, pH 7.5.<br />

16. Wash buffer 2: 0.1 M TRIS, 0.15 M NaCl, pH 9.5.<br />

17. Chromagen: 4-nitro blue tetrazolium chloride (NBT)/5-bromo-4-chloro-3-indoyl phosphate<br />

(BCIP) (Boehringer Mannheim).<br />

18. Substrate solution : 10µL of NBT, 7.5 µL of BCIP in 2 mL of buffer 2.<br />

19. Water, di-ethyl pyrocarbonate (DEPC) treated, RNase free.<br />

20. Silane-coated glass slides.<br />

21. Thermal cycling PCR machine or dedicated in situ PCR machine.<br />

3. Methods<br />

3.1. Cell Fixation<br />

1. Cells collected from tissue culture flasks, blood, or other sources are collected in a 15-mL<br />

polypropylene tube.


448 Gilchrist and Befus<br />

2. Cells are pelleted by centrifugation in a tabletop centrifuge at 800g for 5 min at 4°C.<br />

3. Cells should be washed once by resuspending in 10 mL of PBS and centrifuged again<br />

at 800g for 5 min at 4°C.<br />

4. The supernatant is discarded, and the pelleted cells are then fixed in 10 mL of 10% neutral<br />

buffered formalin for 16 h at room temperature (see Note 2).<br />

5. After fixation the cells are twice washed in DEPC water, recovering the cells each time by<br />

centrifugation at 800g for 5 min at 4°C.<br />

6. The cells are then resuspended in DEPC water.<br />

7. Cells are then dropped (75 µL) onto slides in three discrete areas and allowed to air dry<br />

at room temperature. Cell concentration should be between 5000 to 10,000 cells per<br />

droplet/spot. (Fig. 2) (see Note 3).<br />

3.2. Protease Digestion<br />

1. Dissolve 20 mg of pepsin in 9.5 mL of DEPC-treated water and then add 0.5 mL of 0.2 N<br />

HCl. The final concentration of pepsin should be 5000 U/mL.<br />

2. Add 200 µL of this mix to each spot on the slide (see Note 4). The digestion time has<br />

to be standardized for each cell type (see Note 5). For mast cell lines we use a digestion<br />

time of 45 min at 37°C.<br />

3. After the digestion is completed, the pepsin is inactivated with a 1-min wash in DEPC<br />

water in a Coplin jar.<br />

4. Then, wash the slides in 100% ethanol for 1 min.<br />

5. Allow to air dry at room temperature (see Note 6).<br />

3.3. Pretreatments (Optional)<br />

In our studies of mast cells, the presence of granule-associated heparin in most<br />

preparations is a major obstacle (9). Heparin is a potent inhibitor of RT and Taq<br />

polymerase enzymes, and therefore a heparinase digestion step must be included to<br />

remove heparin before proceeding.<br />

1. 50 µL of heparinase digest buffer containing 333 U/mL of heparinase I is added to the<br />

slide and incubated for 2 h at room temperature.<br />

2. Wash the slides in DEPC water for 1 min.<br />

3. Then wash the slides in 100% ethanol for 1 min.<br />

4. Allow the slides to air dry at room temperature.<br />

3.4. DNase Treatment<br />

This step is used to remove chromosomal DNA. It is added to the negative control<br />

and test cell spots only. Complete removal of genomic DNA is essential for correct<br />

interpretation of the test cells.<br />

1. The negative control and the test cell spots are treated with 50 µL of DNase solution per<br />

spot, and incubated overnight at 37°C.<br />

2. After the overnight digestion, wash the slides in DEPC water for 1 min.<br />

3. Then wash the slides in 100% ethanol for 1 min.<br />

4. Allow the slides to air dry at room temperature.<br />

3.5. Reverse Transcription<br />

1. Add 50 µL of the master mix to the test spot ONLY (see Note 7).<br />

2. Incubate 1 h at 37°C in a moist chamber. Transcripts of low abundance may require longer<br />

incubation of the RT step (see Note 8).


RT In Situ PCR 449<br />

3. Wash slides in DEPC treated water for 1 min.<br />

4. Then wash the slides in 100% ethanol for 1 min.<br />

5. Allow the slides to air dry at room temperature.<br />

3.6. PCR<br />

1. Place 50 µL of the PCR master mix on EACH spot (see Note 9).<br />

2. A good starting program for the TNF primers is outlined: 94°C for 2 min, followed by<br />

30 cycles of: 94°C for 1 min , 45°C for 2 min, 72°C for 2 min, and a final step at 4°C<br />

until ready to develop the slides. However, this may require optimization for each gene<br />

and cell type studied (see Note 10).<br />

3. Once the slides have finished cycling, remove the coverslip and transfer the slides to<br />

wash buffer 1 for 5 min.<br />

4. Do not allow the slides to air dry from this point on.<br />

3.7. Digoxigenin Detection and Color Development<br />

This step uses an alkaline phosphatase (AP)-labeled anti-digoxigenin antibody to<br />

detect the incorporated digoxigenin-dUTP. Color development is accomplished with<br />

NBT/BCIP Chromagen, which is oxidized to a purple/blue color by AP.<br />

1. Prepare a 1300 dilution (see Note 11) of anti-DIG antibody using wash buffer 1 as the<br />

diluent. Add 150 µL of the diluted antibody to the slide. Incubate the slides in a moist<br />

chamber for 30 min at room temperature.<br />

2. Wash the slides with wash buffer 2 for 5 min.<br />

3. Prepare the substrate solution.<br />

4. Add the substrate solution to the slides and develop for 5 min to 1 h at room temperature.<br />

Monitor the purple/blue color development under the microscope.<br />

5. Stop the color reaction by washing the slides in water.<br />

6. Wash the slides for 1 min in 100% ethanol.<br />

7. Wash in xylene, 1 min.<br />

8. Mount using Permount and a coverslip.<br />

3.8. Controls and Expected Results<br />

The positive control (–DNAse, –RT, +PCR) should show intense nuclear staining in<br />

>90% of the cell population. This indicates incorporation of digoxigenin into strand<br />

breaks of genomic DNA by Taq polymerase and ensures that the protease digestion<br />

and PCR were adequate (Fig. 3b).<br />

The negative control (+ DNAse, –RT, +PCR) should show no staining. This indicates<br />

that all genomic DNA has been digested and thus any staining in the test is caused<br />

by mRNA expression (Fig. 3c).<br />

The test (+ DNAse, +RT, +PCR), if positive, should show staining predominantly<br />

over the cytoplasm and not in the nucleus. The cytoplasm is the correct cellular<br />

compartment for most mRNA (Fig. 3a), although it may have an intense perinuclear<br />

localization, especially in highly granulated cells (8).<br />

3.9. Future Directions<br />

Initial development of RT in situ PCR lent itself immediately to areas of pathology<br />

and detection of viruses important in human disease. Publications in the field provided<br />

new insights into the pathogenesis and involvement of viruses, such as human papilloma<br />

virus (HPV) in cervical carcinoma (10) and Human Immunodeficiency Virus (HIV)


450 Gilchrist and Befus<br />

Fig. 3. RT in situ PCR detection of mRNA in a rat mast cell line (RCMC 1.11.2) to show<br />

representative results. (A) Positive signal localized in the cytoplasm of RCMC indicating the<br />

presence of TNF mRNA. (B) Positive control showing nuclear staining only and indicating that<br />

protease digestion, PCR, and detection steps are optimized. (C) Negative control showing no<br />

staining, indicating that genomic DNA is not causing a nonspecific signal. Original magnification<br />

×1000; bar = 10 µm.<br />

(4,11). In both cases, viral nucleic acid was found to be latently expressed in many<br />

cells, a result that was only obtainable with RT in situ PCR.<br />

This methodology is now progressing into the area of cytokine research. Cytokine<br />

production and regulation has been difficult to study because of their low levels of<br />

expression and rapid turnover. Previously, our knowledge of cytokine expression and<br />

interaction was obtained from studies performed in cell culture systems, far removed<br />

from a true in vivo environment. With recent progresses in this method, it has become<br />

feasible to identify the source, kinetics and tissue distribution patterns of cytokines in<br />

situ, and their patterns of distribution in both health and disease (12,13).<br />

The future of RT in situ PCR looks promising. Growth in the area will progress with the<br />

ability to colocalize both mRNA signals by RT in situ PCR, combined with a technique<br />

such as immunohistochemistry for protein detection. Such a protocol will allow researchers<br />

the power to identify both protein and gene expression in the same cell, contributing<br />

greatly to the understanding of molecular interactions at the cellular level.<br />

4. Notes<br />

1. Several methods have been used to confirm the specificity of the products generated by this<br />

methodology (8). In our experience, the labeled PCR product can be directly isolated from<br />

the cells on the slide itself. Once isolated the product can be confirmed either by Southern<br />

blot analysis or by cloning and sequencing of the generated product.<br />

2. Fixation is a critical parameter in this protocol. A wide variety of fixatives have been<br />

used with success by other investigators, including 10% formalin, 4% paraformaldehyde,<br />

ethanol, and methanol/acetic acid (14,15). We have used 10% neutral-buffered formalin as<br />

our fixative of choice because of its good cell structural preservation qualities and minimal<br />

reactivity with RNA (16), thus facilitating good cDNA amplification.<br />

3. Because cells/sections must remain on the slide through the rigors of digestion, pretreatments,<br />

and thermocycling, they must be placed on silane-coated slides. We use slides<br />

available from Perkin–Elmer for our in situ studies. Other sources of slides, including<br />

those produced “in-house” as part of a regular regime for use in immunohistochemistry<br />

can also be used with excellent results.<br />

4. To keep the RT-PCR master mixes from evaporating during cycling, some method of<br />

covering the sections must be devised. We have used several techniques with similar<br />

success, including dedicated neoprene covers produced by Perkin–Elmer, as well as glass


RT In Situ PCR 451<br />

coverslips anchored with nail polish, and polyethylene bags, cut to size and held in place<br />

with nail polish (5,17). It is important to remove ALL visible air bubbles because these will<br />

be expanded as a result of the cycling process, which can result in some parts of sections<br />

being inaccessible to the PCR reagents, resulting in false negatives. Different means of<br />

covering the cell spots will result in different volumes of reagent required for each step,<br />

and they should be adjusted accordingly to save costs.<br />

5. Protease digestion reduces the protein cross-links that form during the fixation step, thus<br />

allowing reagents to enter the cell. We currently use pepsin (5000 U/mL) at 37°C. The<br />

duration of pepsin digestion depends on the type of fixative used and the extent of fixation,<br />

with mast cell, macrophage, and lymphocyte cell lines taking ~45 min of digestion,<br />

whereas other cells such as in vivo derived mast cells requiring much less digestion<br />

(~15 min). However, other temperatures (22°C) and enzymes (trypsin, proteinase K) have<br />

been used by various groups (7). Both these enzymes can give excellent results as well.<br />

We have settled on pepsin as our enzyme of choice, as it works at low pH (3.0), and is thus<br />

easily inactivated by a wash in water (~pH 7.5).<br />

6. As a first step in the process of optimizing the digestion time and to become familiar with<br />

the methodology, we use a shortened protocol (5). This involves using a single slide, with<br />

each individual spot being subjected to a different digestion time (e.g., 20 min, 40 and<br />

60 min). After washing off the protease, the slide is then subjected to PCR and development,<br />

with no DNase or RT steps. By looking at the nuclear staining obtained with the different<br />

times, the investigator will be able to pick the time that gives >90% of the nuclei positive<br />

but maintains the cellular architecture. This optimal time can then be used for complete RT<br />

in situ PCR studies on this same batch of fixed cells. Some investigators have noted that they<br />

use >50% of the nuclei positive as an indication of optimal digestion (4,5). We generally see<br />

>90% positive nuclei and have chosen this as our optimal indicator of digestion.<br />

7. Recommended primer lengths are between 20 to 30 nucleotides. The PCR product should be<br />

at least 300 bp in length to assure that the amplicons remain inside the cell. The RT solution<br />

should contain an antisense primer capable of initiating first strand cDNA production.<br />

Oligo-dT can be used, although we have found that a gene specific primer works best.<br />

8. Numerous RT enzymes are commercially available. We have used both Maloney Murine<br />

Leukemia Virus (MMLV) and Avian Myeloblastosis Virus (AMV) RT enzymes with<br />

identical results. We do our RT steps at 37°C, and vary the time of incubation from 1 to 3<br />

h, depending on the signal seen with a specific cell type.<br />

9. The critical parameter in the composition of the PCR solution is the MgCl 2 concentration.<br />

We found that Mg 2+ concentration in the range of 3.0 to 5.0 mM works best, which is about<br />

four times higher than that used in solution-based PCR. This necessary increase in Mg 2+<br />

concentration is thought to be caused by binding of Mg 2+ to the glass slides (5,7).<br />

10. If no product is detected with the cycling program given, then the program will have to<br />

be optimized. Take into account the T m of the primers being used, and vary the number of<br />

cycles. In general terms, we begin with an annealing temperature that is 5°C below that of<br />

the primer Tm. Furthermore, depending on the abundance of the mRNA being amplified,<br />

upwards of 35 cycles may be required to obtain a detectable signal.<br />

11. Transposed onto the technique of in situ RT-PCR is the necessity to detect the amplified<br />

product using immunohistochemical techniques. This introduces numerous other variables<br />

into the protocol. We have found that antibody concentrations of 0.75 to 3.75 µg/mL<br />

provide optimal signal with little background staining.<br />

Acknowledgments<br />

The authors recognize Dr. Osamu Nohara for his contribution to the development of<br />

our understanding of this procedure.


452 Gilchrist and Befus<br />

References<br />

1. Mullis, K. B. and Faloona, F. A. (1987) Specific synthesis of DNA in vitro via a polymerasecatalysed<br />

chain reaction. Methods Enzymol. 155, 335–350.<br />

2. Veres, G., Gibbs, R. A., Scherer, S. E., and Caskey, C. T. (1987) The molecular basis of the<br />

sparse fur mouse mutation. Science 237, 415– 417.<br />

3. Long, A. A. (1998) In situ polymerase chain reaction: foundation of the technology and<br />

today’s options. Eur. J. Histochem. 42, 101–109.<br />

4. Nuovo, G. J., Forde, A., MacConnell, P., and Fahrenwald, R. (1993) In situ detection of<br />

PCR-amplified HIV-1 nucleic acids and tumor necrosis factor cDNA in cervical tissues.<br />

Am. J. Pathol. 143, 40– 48.<br />

5. Nuovo, G. J. (1992) PCR in situ Hybridization: Protocols and Applications. Raven Press,<br />

New York.<br />

6. Chen, R. H. and Fuggle, S. V. (1993) In situ cDNA polymerase chain reaction. Am.<br />

J. Pathol. 143, 1527–1534.<br />

7. Teo, I. A. and Shaunak, S. (1995) Polymerase chain reaction in situ: An appraisal of an<br />

emerging technique. Histochem. J. 27, 647–659.<br />

8. Nohara, O., Gilchrist, M., Dery, R. E., Stenton, G. R., Hirji, N. S., and Befus, A. D. (1999)<br />

Reverse transcriptase in situ polymerase chain reaction for gene expression in rat mast cells<br />

and macrophages. J. Immunol. Methods 226, 147–158.<br />

9. Gilchrist, M., MacDonald, A. J., Neverova, I., Ritchie, B., and Befus, A. D. (1997)<br />

Optimization of the isolation and effective use of mRNA from rat mast cells. J. Immunol.<br />

Methods 201, 207–214.<br />

10. Chiu, K. P., Cohen, S. H., Morris, D. W., and Jordan, G. W. (1992) Intracellular amplification<br />

of proviral DNA in tissue sections using the polymerase chain reaction. J. Histochem.<br />

Cytochem. 40, 333–341.<br />

11. Patterson, B. K., Till, M., Otto, P., Goolsby, C., Furtado, M. R., McBride, L. J., et al. (1993)<br />

Detection of HIV-1 DNA and messenger RNA in individual cells by PCR-driven in situ<br />

hybridization and flow cytometry. Science 230, 1350–1354.<br />

12. Hagimoto, N., Kuwano, K., Nomoto, Y., Kunitake, R., and Hara, N. (1997) Apoptosis and<br />

expression of Fas/Fas ligand mRNA in bleomycin-induced pulmonary fibrosis in mice. Am.<br />

J. Respir. Cell Mol. <strong>Bio</strong>l. 16, 91–101.<br />

13. Peters, J., Krams, M., Wacker, H. H., Carstens, A., Weisner, D., Hamann, K., et al. (1997)<br />

Detection of rare RNA sequences by single-enzyme in situ reverse transcription polymerase<br />

chain reaction: High-resolution analysis of interleukin-6 mRNA in paraffin sections of<br />

lymph nodes. Am. J. Pathol. 150, 469– 476.<br />

14. O’Leary, J. J. (1994) Importance of fixation procedures on DNA template and its suitability<br />

for solution phase PCR and PCR in situ hybridization. Histochem. J. 26, 337–346.<br />

15. Long, A. A., Komminoth, P., and Wolf, H. J. (1993) Comparison of indirect and direct<br />

in situ polymerase chain reaction in cell preparations and tissue sections. Histochemistry<br />

99, 151–162.<br />

16. Teo, I. A. and Shaunak, S. (1995) PCR in situ: aspects which reduce amplification and<br />

generate false-positive results. Histochem. J. 27, 660–669.<br />

17. Patel, V. G., Shum-Siu, A., Heniford, B. W., Weiman, T. J., and Hendler, F. J. (1994)<br />

Detection of epidermal growth factor receptor mRNA in tissue sections from biopsy<br />

specimens using in situ polymerase chain reaction. Am. J. Pathol. 144, 7–14.


PRINS and Immunocytochemistry 453<br />

63<br />

Primed In Situ Nucleic Acid Labeling Combined with<br />

Immunocytochemistry to Simultaneously Localize DNA<br />

and Proteins in Cells and Chromosomes<br />

Ernst J. M. Speel, Frans C. S. Ramaekers, and Anton H. N. Hopman<br />

1. Introduction<br />

In the past decade, the primed in situ (PRINS) labeling technique has become an<br />

alternative to fluorescence in situ hybridization (ISH) for the localization of nucleic acid<br />

sequences in chromosome, cell, and tissue preparations (1–8). The PRINS method is<br />

based on the rapid annealing of unlabeled primers (restriction fragment, PCR product,<br />

or oligonucleotide) to complementary target sequences in situ. These primers serve<br />

as initiation sites for in situ chain elongation using Taq DNA polymerase and labeled<br />

nucleotides. Incorporated fluorochrome-labeled nucleotides can be detected directly by<br />

fluorescence microscopy, whereas haptenized (e.g., biotin, digoxigenin, dinitrophenyl)<br />

nucleotides can be visualized by the additional application of fluorochrome- or<br />

enzyme-conjugated avidin or antibody molecules (4,5,9,10), followed by fluorescence<br />

microscopy or brightfield visualization of enzyme reaction products. Particularly,<br />

rapidity, simplicity, and cost-effectiveness have made the PRINS technique a useful<br />

tool in cytogenetics and cell biology. Its detection sensitivity, however, seems to<br />

be limited to repetitive targets for a long time. Only recently, the combined use of<br />

multiple oligonucleotides for 1 locus together with tyramide signal amplification<br />

have shown the first reproducible results demonstrating single-copy gene detection<br />

by PRINS (11).<br />

With immunocytochemistry (ICC), specific <strong>info</strong>rmation can be obtained regarding<br />

the presence or absence of proteins or antigens in chromosomes, cells, and tissue<br />

sections, thus allowing one to characterize the function of structural proteins in<br />

chromosomes or to phenotype cells (for example, their type of differentiation and<br />

proliferative activity). In addition, protocols have been developed to efficiently combine<br />

protein and nucleic acid detection in the same biological material to, for example,<br />

immunophenotype cells harboring a specific chromosomal aberration or viral infection,<br />

determine cytokinetic parameters of tumor cell populations that are genetically of<br />

phenotypically aberrant, and study the structural organization of chromosomes and the<br />

cell nucleus (10,12). The success and sensitivity of such a combined procedure depends<br />

on factors, such as preservation of cell morphology and protein epitopes, accessibility<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

453


454 Speel, Ramaekers, and Hopman<br />

of nucleic acid targets, lack of crossreaction between the different protein and nucleic<br />

acid detection procedures, and good color contrast and stability of the fluorochromes<br />

and enzyme cytochemical precipitates applied. A variety of procedures have been<br />

reported that combine ICC and ISH (for a review, see ref. 10,12), in most cases<br />

applying ICC before ISH to prevent the destruction of antigenic determinants by<br />

the ISH procedure because of enzymatic digestion, postfixation, denaturation at<br />

high temperatures, and hybridization in formamide. For reasons of rapidity, probe<br />

accessibility, and lack of formamide for hybridization, the substitution of ISH by<br />

PRINS may be an extra advantage in such a combined procedure.<br />

Here, we present three protocols for combined ICC and PRINS DNA labeling. In the<br />

first procedure, a sensitive, high-resolution fluorescence alkaline phosphatase (APase)-<br />

Fast Red ICC staining method (13,14) is performed before subsequent PRINS labeling<br />

of DNA target sequences to enable the simultaneous detection of surface antigens<br />

(EGF receptor, neural cell adhesion molecule) and repeated chromosome-specific DNA<br />

sequences in somatic cell hybrid and tumor cell lines. The fact that the Apase-Fast Red<br />

precipitate withstand subsequent proteolytic digestion and denaturation steps guarantees<br />

the most optimal conditions for efficient PRINS labeling (15). The second procedure<br />

has been recently described to investigate chromosome distribution and segregation in<br />

cells during processes, such as polyploidization and aneuploidization. This protocol<br />

starts with the PRINS labeling of chromosome centromeres followed by the staining<br />

of the mitotic spindle by tubulin ICC, because the authors found the antibody epitope<br />

to be better preserved when ICC came after DNA labeling (16,17). The third protocol<br />

has been used on metaphase chromosomes to identify possible relationships of different<br />

families of DNA sequences with, for example, proteins associated with different<br />

chromosome-specific structures, such as the kinetochore complex. This approach again<br />

starts with ICC staining followed by PRINS DNA labeling (18,19).<br />

2. Materials<br />

2.1. Protocol 1<br />

2.1.1. Fluorescence Immunophenotyping by Alkaline Phosphatase Cytochemistry<br />

1. Cold methanol (–20°C), cold acetone (4 and –20°C), and cold 70% ethanol (–20°C).<br />

2. Normal goat serum (NGS).<br />

3. Monoclonal antibody EGFR1, directed against the epidermal growth factor receptor (a<br />

kind gift of V. van Heyningen, Edinburgh, UK).<br />

4. Monoclonal antibody 163A5, directed against a cell-surface marker of J1-C14 cells (20).<br />

5. Monoclonal antibody RNL1, directed against the neural cell adhesion molecule<br />

(N-CAM) (21).<br />

6. Alkaline phosphatase-conjugated goat anti-mouse IgG (GAMAPase) (Dako, Glostrup,<br />

Denmark).<br />

7. Naphthol-ASMX-phosphate (Sigma, St. Louis, MO).<br />

8. Fast Red TR (Sigma).<br />

9. Polyvinylalcohol (PVA), MW 40,000 (Sigma).<br />

10. 10× phosphate-buffered saline (PBS): 1.37 M NaCl, 30 mM KH 2 PO 4 , 130 mM Na 2 HPO 4 .<br />

11. APase buffer: 0.2 M Tris-HCl, pH 8.5, 10 mM MgCl 2 , 5% PVA. Dilute from stock solutions<br />

1 M Tris-HCl, pH 8.5 and 1 M MgCl 2 and add 5% (w/v) PVA. Dissolve PVA by using<br />

a microwave.<br />

12. Blocking buffer: 1× PBS (diluted from stock 10× PBS), 0.05% Triton X-100, 2–5% NGS.


PRINS and Immunocytochemistry 455<br />

Table 1<br />

Sequences of Oligonucleotide Primers Used in PRINS<br />

Name Human origin DNA sequence<br />

E528 Chromosome 7 centromere AGCGATTTGAGGACAATTGC<br />

G33 Chromosome 9 centromere AATCAACCCGAGTGCAATC<br />

G35 Chromosome 11 centromere GAGGGTTTCAGAGCTGCTC<br />

D600 Chromosome X centromere TCCATTCGATTCCATTTTTTTCGAGAA<br />

13. Washing buffer: 1× PBS, 0.05% Triton X-100.<br />

14. Glass coverslips 50 × 24 mm (Menzel Gläzer, Braunschweig, Germany).<br />

15. Humid chamber.<br />

16. Coplin jars (50 or 100 mL).<br />

2.1.2 PRINS DNA Labeling<br />

1. Pepsin from porcine stomach mucosa (2500–3500 U/mg; Sigma).<br />

2. Ultrapure dNTP set (Amaersham Pharmacia <strong>Bio</strong>tech, Little Chalfont, UK): 100 mM<br />

solutions of dATP, dCTP, dGTP, and dTTP.<br />

3. <strong>Bio</strong>tin-16-dUTP, Digoxigenin-11-dUTP, and Fluorescein-12-dUTP (Roche Molecular<br />

<strong>Bio</strong>chemicals, Basel, Switzerland).<br />

4. Oligonucleotide primer (see Table 1 and Note 1).<br />

5. Taq 1 DNA polymerase (Roche) or AmpliTaq (Perkin–Elmer).<br />

6. Bovine serum albumin (BSA; Sigma).<br />

7. Dried skimmed milk powder.<br />

8. FITC-conjugated avidin (AvFITC; Vector, Burlingame, CA).<br />

9. FITC-conjugated sheep anti-digoxigenin Fab fragments (SHADigFITC; Roche).<br />

10. Vectashield (Vector).<br />

11. 4′,6-diamidino-2-phenyl indole (DAPI; Sigma).<br />

12. 0.01 M HCl.<br />

13. 1× PBS: diluted from 10× PBS (see Subheading 2.1.1., item 10).<br />

14. Postfixation buffer: 1% formaldehyde (diluted from 37% formaldehyde (Merck, Darmstadt,<br />

Germany) in 1× PBS.<br />

15. 20 × SSC: 3 M NaCl, 300 mM trisodium citrate, pH 7.0.<br />

16. 10× Taq buffer: 500 mM KCl; 100 mM Tris-HCl, pH 8.3, 15 mM MgCl 2 , 0.1% BSA.<br />

17. PRINS stop buffer: 500 mM NaCl, 50 mM EDTA, pH 8.0.<br />

18. Blocking buffer: 4 × SSC (diluted from stock 20 × SSC), 0.05% Triton X-100, 5%<br />

skimmed milk powder.<br />

19. Washing buffer: 4 × SSC, 0.05% Triton X-100.<br />

20. Coplin jars (50 or 100 mL).<br />

21. Ethanol/37% HCl (1001)-cleaned microscope slides and coverslips (Menzel).<br />

22. Rubber cement.<br />

23. Water bath at 65°C.<br />

24. Thermal cycler (Hybaid Omnigene Flatbed; Hybaid, Ashford, UK; see Note 2).<br />

25. Humid chamber.<br />

26. Incubator set at 37°C.<br />

27. Leica DM fluorescence microscope (Leica, Wetzlar, Germany) equipped with filter sets<br />

for DAPI, FITC, and TRITC.<br />

28. Kodak 400 ASA film.<br />

29. Black and white CCD camera and Isis 4 analysis software (Metasystems, Sandhausen,<br />

Germany).


456 Speel, Ramaekers, and Hopman<br />

Fig. 1. Fluorescence detection of (A) chromosome 9 centromeres with digoxigenin/<br />

SHADigFITC in T24 cells after PRINS and Vectashield embedding with PI counterstaining. (B)<br />

chromosome 7 and 9 centromeres with, respectively, biotin/AvidinTexasRed anf digoxigenin/<br />

SHADigFITC after double-target PRINS and Vectashield embedding with DAPI counterstaining.<br />

(C) EGF receptor with alkaline phosphatase-Fast Red and four chromosome 7 centromeres<br />

with biotin/avidinFITC in C121-TN6 cells after immunocytochemistry followed by PRINS<br />

and Vectashield embedding. (D) NCAM antigen with alkaline phosphatase-Fast Red and three<br />

chromosome 9 centromeres with digoxigenein/SHADigFITC in H460 cells after immunocytochemistry<br />

followed by PRINS and Vectashield embedding.<br />

30. <strong>Bio</strong>–Rad MRC 600 confocal scanning laser microsope equipped with the laser lines 488,<br />

568, and 648 nm (<strong>Bio</strong>–Rad Laboratories, Veenendaal, The Netherlands).<br />

2.2. Protocol 2<br />

2.2.1. PRINS DNA Labeling<br />

1. Dried skimmed milk powder.<br />

2. PRINS reaction mixture components: dNTPs, labeled dUTPs, oligonucleotide primers, and<br />

Taq DNA polymerases as described in Subheading 2.1.2., items 2–6 and 16.<br />

3. 1× PBS (diluted from 10× PBS, see Subheading 2.1.1., ref. 10) containing 0.1% EDTA<br />

and 0.2% BSA.<br />

4. Methanol: acetic acid (91), freshly prepared.<br />

5. 1× PBS (diluted from 10× PBS, see Subheading 2.1.1., ref. 10).<br />

6. Permeabilization buffer: 1× PBS containing 0.1% Triton X 100.<br />

7. A series of 70, 96, and 100% ethanol.<br />

8. Denaturation buffer: 70% formamide/2 × SSC pH 7.0.


PRINS and Immunocytochemistry 457<br />

9. Washing buffer: 4 × SSC (diluted from 20 × SSC, see Subheading 2.1.2., item 15),<br />

0.05% Triton X 100.<br />

10. Blocking buffer: 4 × SSC (diluted from 20 × SSC, see Subheading 2.1.2., item 15), 0.05%<br />

Triton X 100, 5% dried skimmed milk powder.<br />

11. Glass slides (Menzel).<br />

12. Coplin jars.<br />

13. Cytocentrifuge (Shandon, Astmoor, UK).<br />

14. Waterbath at 70°C.<br />

15. Thermal cycler (Hybaid Omnigene Flatbed; Hybaid).<br />

16. Incubator at 37°C.<br />

2.2.2. Immunofluorescence Detection of Mitotic Spindle<br />

1. Mouse monoclonal anti-β-tubulin antibody (Sigma).<br />

2. TRITC-conjugated donkey anti-mouse IgG F(ab′)2 fragments (DAMTRITC; Jackson<br />

Immunoresearch, West Grove, PE).<br />

3. Normal goat serum (NGS).<br />

4. Vectashield (Vector).<br />

5. DAPI (see Subheading 2.1.2., item 11).<br />

6. TOTO-3 iodide (Molecular Probes, Eugene, OR).<br />

7. Blocking buffer: 1× PBS (diluted from stock 10× PBS, see Subheading 2.1.1., item 9),<br />

0.05% Triton X-100, 2-5% NGS.<br />

8. Washing buffer: 1× PBS (diluted from stock 10× PBS, see Subheading 2.1.1., item 9),<br />

0.05% Triton X-100.<br />

9. 1× PBS.<br />

10. Vectashield (Vector) containing 0.5 µg/mL DAPI or 3000× diluted TOTO-3 iodide (from<br />

1 mM stock in DMSO).<br />

11. Humid chamber.<br />

12. Coplin jar (50 or 100 mL).<br />

13. Fluorescence microscope, CCD camera, and confocal microscope (see Subheading 2.1.2.,<br />

items 27–30).<br />

2.3. Protocol 3<br />

2.3.1. Immunofluorescence Detection of Structural Proteins in Chromosomes<br />

1. Normal goat serum (NGS).<br />

2. Primary antibody.<br />

3. FITC-conjugated secondary antibody.<br />

4. Alcohol-cleaned glass slides and coverslips (Menzel).<br />

5. Hypotonic solution: 75 mM KCl.<br />

6. Potassium chromosome medium (KCM) solution: 120 mM KCl, 20 mM NaCl, 10 mM<br />

Tris-HCl, pH 8.0, 0.5 mM EDTA, and 0.1% (v/v) Triton X-100.<br />

7. Blocking buffer: KCM containing 2 to 5% NGS.<br />

8. Fixation solution: KCM containing 10% formalin (from 37% stock solution; Merck).<br />

9. Cytocentrifuge (Shandon).<br />

10. Coplin jar (50 or 100 mL).<br />

11. Humid chamber.<br />

2.3.2. PRINS DNA Labeling<br />

1. Immersion solution: 0.1 M NaOH.<br />

2. Neutralization solution: 0.01 M Tris-HCl, pH 7.4.


458 Speel, Ramaekers, and Hopman<br />

3. Methanol:acetic acid (31).<br />

4. A series of cold (4°C) 70, 96, and 100% ethanol.<br />

5. Denaturation solution: 30 mM NaOH/1 M NaCl (pH >12).<br />

6. PRINS reaction, detection, and mounting components, as described in Subheading 2.1.2.,<br />

items 2–26.<br />

7. Slide evaluation equipment, as described in Subheading 2.1.2., items 27–30.<br />

3. Methods<br />

3.1. Protocol 1<br />

3.1.1. Fluorescence Immunophenotyping by Alkaline Phosphatase Cytochemistry<br />

1. Hybrid (C121-TN6, J1-C14) and tumor (H460) cell lines are cultured on glass slides<br />

by standard methods (20,22,23), fixed in either cold methanol (–20°C) for 5 s and cold<br />

acetone (4°C) for 3 × 5 s, cold acetone (–20°C) for 10 min, or 70% ethanol (–20°C) for<br />

10 min, air-dried, and stored at –20°C until use (see Note 3).<br />

2. Incubate slides for 10 min at room temperature with 100 µL of blocking buffer under a<br />

coverslip in a humid chamber.<br />

3. Remove the coverslip, discard the blocking buffer, and incubate the slides for 30 to 45 min<br />

at room temperature with 50 µL of undiluted culture supernatant of the appropriate antigenspecific<br />

monoclonal antibody containing 2% NGS under a coverslip in a humid chamber.<br />

4. Wash slides for 2 × 5 min with washing buffer in a Coplin jar.<br />

5. Incubate slides for 30 to 45 min at room temperature with 50 µL of GAMAPase, diluted<br />

150 in blocking buffer (see Note 4), under a coverslip in a humid chamber.<br />

6. Wash slides for 5 min with washing buffer, and for 5 min with 1× PBS.<br />

7. Visualize the antigen with the alkaline phosphatase-Fast Red (APase-Fast Red) reaction:<br />

Mix 4 mL of APase buffer, 1 mg of naphthol-ASMX-phosphate in 250 µL of buffer without<br />

PVA and 5 mg of Fast Red TR in 750 µL of buffer without PVA just before use and overlay<br />

each sample with 100 µL under a coverslip. Incubate the slides for 5 to 15 min at 37°C<br />

and wash 3× 5 min with 1× PBS (see Notes 5–7).<br />

3.1.2 PRINS DNA Labeling<br />

1. Process cells for PRINS as follows: wash slides for 2 min at 37°C with 0.01 M HCl, incubate<br />

the samples with 100 µg/mLpepsin in 0.01 M HCl for 20 min at 37°C, wash again with<br />

0.01 M HCl for 2 min, and post-fix the slides in 1% formaldehyde in 1× PBS for 10 min<br />

at room temperature. Wash cells in 1× PBS for 5 min at room temperature, followed by a<br />

wash step in 1× Taq buffer for 5 min at room temperature (all steps in a Coplin jar).<br />

2. Prepare the PRINS reaction mix on ice as follows: Dilute 100 mM dATP, dGTP, and dCTP<br />

110 with distilled water. Dilute 100 mM dTTP 1:100. Put together in a microcentriphuge<br />

tube: 1 µL of each of the diluted dNTPs, 1 µL of either 1mM <strong>Bio</strong>tin-16-dUTP, Digoxigenin-<br />

11-dUTP, or Fluorescein-12-dUTP (see Note 8), 5 µL of 10× Taq buffer, 250 ng of<br />

oligonucleotide (see Note 9), 1 U Taq polymerase, and distilled water to 50 µL.<br />

3. Place 40 µL of this mixture under a coverslip on the slide, seal with rubber cement, air-dry<br />

the rubber cement, and transfer to the heating block of the thermal cycler.<br />

4. Each PRINS reaction cycle consists of 2 min at 94°C (denaturation of cellular DNA, see<br />

Note 10), 5 min at the appropriate annealing temperature (see Note 11) and 15 min at<br />

72°C for in situ primer extension.<br />

5. Stop the PRINS reaction by transferring the slides (after removal of the rubber solution<br />

seal) to 50 mLof PRINS stop buffer in a Coplin jar at 65°C for 1 min (see Note 12).<br />

6. Transfer the slides to washing buffer at room temperature and wash 5 min.


PRINS and Immunocytochemistry 459<br />

7. Place 40 µL of blocking buffer under a coverslip on the slide and leave for 5 min at room<br />

temperature in a humid chamber to reduce background staining in the detection procedures.<br />

8. Wash slides for 1 × 5 min in washing buffer in a Coplin jar.<br />

9a. For reactions using <strong>Bio</strong>tin-16-dUTP: Dilute AvFITC 1:100 in blocking buffer and apply<br />

50 µL under a coverslip. Incubate slides for 30 min at 37°C in a humid chamber (see<br />

Note 13).<br />

9b. For reactions using Digoxigenin-11-dUTP: Dilute SHADigFITC 1100 in blocking buffer<br />

and treat as in 9a (see Note 13).<br />

9c. Fluorescein-12-dUTP needs no additional reporter and is simply mounted as described<br />

in 11 (see Note 13).<br />

10. Wash slides for 2 × 5 min in washing buffer in a Coplin jar. Optionally, you may wash the<br />

slides for 5 min in 1× PBS and dehydrate them.<br />

11. Mount the slides in Vectashield containing 0.5 µg/mLDAPI.<br />

12. Examine the slides under a fluorescence microscope. Selected cells can be either directly<br />

photographed using Kodak 400 ASA film, visualized with a charge-coupled device (CCD)<br />

camera, or scanned with a confocal scanning laser microscope (CSLM).<br />

3.2. Protocol 2<br />

3.2.1 PRINS DNA Labeling<br />

1. Hybrid and tumor cell lines are cultured on glass slides, as described in Subheading<br />

3.1.1., step 1), or in suspension according to standard methods (16,17,20,22,23). Cells are<br />

rinsed in 1× PBS/0.1% EDTA/0.2% BSA, cytocentrifuged on glass slides at 65g for 4 min<br />

(in case of growth in suspension), fixed in methanolacetic acid (91) for 10 min at room<br />

temperature in a Coplin jar, and airdried.<br />

2. Rehydrate cells in 1× PBS for 5 min, permeabilize them with 0.1% Triton X 100 in<br />

1× PBS for 5 to 10 min (both in a Coplin jar), dehydrate in a series of 70, 96, and 100%<br />

ethanol and airdry.<br />

3. Denature slides in 70% formamide/2 × SSC, pH 7.0, for 2 min at 70°C in a Coplin jar,<br />

dehydrate in a series of 70, 96 (both at 4°C), and 100% ethanol, and airdry.<br />

4. Prepare the PRINS reaction mix and apply it on the prewarmed (annealing temperature)<br />

slides (on the block of the thermal cycler) as described in Subheading 3.1.2., steps 2<br />

and 3).<br />

5. Perform the PRINS reaction for 5 min at the appropriate annealing temperature (see Note 11)<br />

and for 15 min at 72°C for chain elongation on a thermal cycler.<br />

6. Stop the PRINS reaction by removing the coverslips (see Note 12) and wash the slides<br />

in washing buffer for 3 × 5 min at room temperature followed by a 5-min wash step in<br />

1× PBS (all steps in a Coplin jar).<br />

7. Apply a blocking and wash step as described in Subheading 3.1.2., steps 7 and 8.<br />

8. In case of PRINS labeling with <strong>Bio</strong>tin-16-dUTP or Digoxigenin-11-dUTP, detect the<br />

haptens as described in Subheading 3.1.2., steps 9 and 10).<br />

3.2.2. Immunofluorescence Detection of the Mitotic Spindle<br />

1. Incubate the slides for 30 to 45 min at room temperature with 50 µL of mouse<br />

anti-β-tubulin primary antibody, diluted 150 in blocking buffer, under a coverslip in<br />

a humid chamber.<br />

2. Wash the slides for 2 × 5 min with washing buffer in a Coplin jar.<br />

3. Incubate slides for 30 to 45 min at room temperature with 50 µL of DAMTRITC, diluted<br />

1100 in blocking buffer (see Note 4), under a coverslip in a humid chamber.


460 Speel, Ramaekers, and Hopman<br />

4. Wash the slides for 2 × 5 min with washing buffer and for 5 min with 1× PBS (all steps<br />

in a Coplin jar).<br />

5. Mount slides in Vectashield containing 0.5 µg/mLDAPI or 3000× diluted TOTO-3 iodide.<br />

6. Examine the slides as described in Subheading 3.1.2., step 12.<br />

3.3. Protocol 3<br />

3.3.1. Immunofluorescence Detection of Structural Proteins in Chromosomes<br />

1. Prepare metaphase spreads from human peripheral blood lymphocytes as described<br />

previously (18,19).<br />

2. After hypotonic treatment of the cell suspension for 10 min at 37°C in 75 mM KCl,<br />

cytocentrifuge the cells onto alcohol-cleaned glass slides with 65–275g for 4 min and<br />

air dry for 2 min.<br />

3. Place slides in a Coplin jar containing KCM solution for 10 min at room temperature,<br />

followed by incubation for 10 min at room temperature with 100 µL of blocking buffer<br />

under a coverslip in a humid chamber.<br />

4. Incubate slides for 30 to 45 min at room temperature with 50 µL of the primary antibody,<br />

diluted in blocking buffer, under a coverslip in a humid chamber.<br />

5. Wash slides for 2 × 5 min with KCM in a Coplin jar.<br />

6. Incubate slides for 30 to 45 min at room temperature with 50 µL of FITC-conjugated<br />

secondary antibody, diluted in blocking buffer, under a coverslip in a humid chamber<br />

(see Note 4).<br />

7. Wash slides for 2 × 5 min with KCM, and fix the chromosomes in fixation solution for 5<br />

to 15 min at room temperature in a Coplin jar.<br />

8. Wash slides in destilled water, airdry and store in the dark at room temperature until use.<br />

3.2.3. PRINS DNA Labeling<br />

1. To improve the signal intensity after PRINS labeling, the slides are immersed in 0.1 M NaOH<br />

for 10 to 40 s (see Note 14) followed by neutralization with 2 × 5-min washes in 0.01 M<br />

Tris-HCl, pH 7.4, and a wash in destilled water (all steps in a Coplin jar) before airdrying.<br />

2. Fix slides in methanolacetic acid (31) for 2 × 2 min in a Coplin jar (see Note 15), wash<br />

in 0.01 M Tris-HCl, pH 7.4, for 2 × 5 min in a Coplin jar, dehydrate slides in a series of<br />

cold (4°C) 70, 96, and 100% ethanol, and airdry (see Note 16).<br />

3. Denature chromosomal DNA in denaturation solution for 45 min at 4°C and neutralize in<br />

0.01 M Tris-HCl, pH 7.4, for 2 × 5 min at room temperature (all steps in a Coplin jar).<br />

4. Remove excess fluid by draining and blow slides dry with a jet of air.<br />

5. Perform PRINS reaction, mounting and evaluation of the slides as described in Subheading<br />

3.1.2., steps 2–12).<br />

4. Notes<br />

1. So far, specific oligonucleotide primers have been defined for the centromere regions<br />

of up to 20 chromosomes (7,24,25). Particularly, the ability of primers to differentiate<br />

between closely related sequences has made it possible to define α-satellite primers<br />

for some chromosomes indistinguishable by FISH with centromeric probes, such as<br />

for chromosomes 13 and 21, that only exhibit a one-base differerence at their 3′ end.<br />

Furthermore, a number of chromosome-specific telomere primers have been generated, as<br />

well as HPV-specific and single-gene-specific oligonucleotides (11,26,27).<br />

2. Several companies commercialize specialized thermal cyclers with a flat block, including<br />

Hybaid, Perkin–Elmer, Techne Corporation (Cambridge, UK), and MJ Research Inc.<br />

(Watertown, MA). Because of differences in the design of the heating block, PRINS


PRINS and Immunocytochemistry 461<br />

conditions always need to be optimized for the respective instrument used. Because an<br />

accurate temperature at the top surface of the slide is crucial for a successful PRINS<br />

reaction, some cyclers, such as the Hybaid Omnigene, possess incorporated software<br />

to compensate for the temperature difference between the block and the surface of the<br />

slide.<br />

3. Because on methanol-acetone fixed cells we frequently observed a poor preservation of<br />

cell morphology as well as fluorescent staining of the entire nucleus (by PRINS labeling),<br />

probably caused by nuclease activities that survive this mild type of fixation, other fixatives<br />

should be and have been successfully tested that are compatible with antigen detection and<br />

result in a better cell morphology and specific PRINS labeling. For example, a fixation with<br />

cold 70% ethanol (–20°C) for 10 min proved to be a valid alternative on H460 cells.<br />

4. If further amplification of the (immuno)cytochemical signal is needed, a third detection<br />

step may be added after this second incubation step. Alternatively, the avidin biotinylated<br />

enzyme (e.g., alkaline phosphatase) complex system may be applied as well as the tyramide<br />

signal amplification system (in combination with peroxidase conjugates). For details of<br />

possible reagents to use, see Note 13 (7,10,12).<br />

5. It is recommended to monitor the enzyme reaction under the microscope to adjust the<br />

reaction time to ensure the precipitate becoming discretely localized and not so dense that<br />

it shields nucleic acid sequences in the PRINS reaction.<br />

6. To ensure the specificity of the APase-Fast Red staining, a control slide with FITCconjugated<br />

secondary antibodies is recommended for comparison. Staining specificity<br />

can be lost if cells contain endogenous APase activity. This endogenous enzyme activity<br />

can be inhibited by the addition of levamisole (Sigma) to the reaction medium to a final<br />

concentration of 1–5 mM.<br />

7. Do not dehydrate the slides after the APase reaction because the precipitate dissolves in<br />

organic solvents. Optionally, you may air dry the slides after rinsing in distilled water.<br />

8. In the case of labeling with biotin-16-dUTP or fluorescein-12-dUTP a 4× decrease of<br />

the concentration of dTTP in the PRINS reaction mix resulted in significant stronger<br />

labeling of DNA sequences. Under the described standard conditions, digoxigenin-<br />

11-dUTP provides the highest sensitivity. However, all the modified nucleotides are suited<br />

for detection of one or multiple repeated sequences in situ. In this respect, the number of<br />

different fluorochrome- and hapten-labeled nucleotides is still increasing to date, which<br />

can be obtained from a number of companies, such as, e.g., Perkin–Elmer Life Science,<br />

Roche, Amersham, Dako, and Molecular Probes.<br />

9. The concentration of the appropriate oligonucleotide resulting in positive signals need to<br />

be determined by experiment. Generally, 250 ng/slide (50–250 pmol) in 40 µL is used for<br />

primers of 16 to 35 bases complementary to repeated sequences.<br />

10. Seperate denaturation of cellular DNA in 70% formamide/2xSSC, pH 7.0, for 2 min at 70°C<br />

before the PRINS reaction as is usually performed for chromosome preparations, resulted<br />

in no or only weak PRINS labeling of DNA sequences in the interphase nuclei of the cell<br />

preparations. Whether this is caused by inefficient primer annealing or extension is not clear.<br />

The same phenomenon is also observed for PRINS on frozen tissue sections (6,7).<br />

11. The optimum primer annealing temperature is only determined empirically. We usually<br />

try a series from 45 to 70°C in 5°C steps.<br />

12. For detection of multiple DNA targets by sequential PRINS reactions (MULTIPRINS), it<br />

was found essential to prevent the free 3′ ends of the newly synthesized DNA from being<br />

used as primers for subsequent reactions. This can be achieved by incubating the slides<br />

with Klenow DNA polymerase together with ddNTPs. The reaction mix is made up as<br />

follows: Dilute 5 mM of all four ddNTPs 110 with distilled water. Put together in a<br />

centrifuge tube 2.5 µL of each of the ddNTPs (Amersham Pharmacia), 5 µL of 10× Klenow


462 Speel, Ramaekers, and Hopman<br />

buffer (500 mM Tris-HCl, pH 7.2; 100 mM MgSO 4 ; 100 mM DTT; 1.5 mg.mLBSA), 1 U<br />

Klenow DNA polymerase (Roche), and distilled water to 50 µL. Incubate slide with<br />

40 µL under a coverslip for 1 h at 37°C in a humid chamber, followed by dehydration and<br />

airdrying before running the next PRINS reaction (5).<br />

13. Amplification of PRINS signals can be achieved as follows:<br />

a. AvFITC detection of <strong>Bio</strong>tin-16-dUTP may be followed by incubation with biotinylated<br />

goat anti-avidin (Vector), 1100 diluted in blocking buffer, and again AvFITC.<br />

b. SHADigFITC detection of digoxigenin-11-dUTP may be followed by incubation<br />

with FITC-conjugated anti-sheep IgG (Roche), or as described for FITC-12-dUTP<br />

amplification.<br />

c. Fluorescein-12-dUTP signals may be amplified by incubation with monoclonal mouse<br />

or polyclonal rabbit anti-FITC (Dako) and FITC-conjugated rabbit anti-mouse IgG or<br />

swine anti-rabbit IgG (Dako).<br />

d. Amplification of PRINS signals may also be achieved by using peroxidase-mediated<br />

deposition of hapten- or fluorochrome-labeled tyramides (28–30).<br />

14. The time required will vary according to the repeated DNA family of sequences under study<br />

(size of target) and the cell type from which the chromosome preparations are made.<br />

15. This again may vary according to the cell type and target size under study.<br />

16. The slides can be stored in the dark for several weeks at room temperature at this stage.<br />

References<br />

1. Koch, J., Kölvraa, S., Petersen, K. B., Gregersen, N., and Bolund, L. (1989) Oligonucleotidepriming<br />

methods for the chromosome-specific labelling of alpha satellite DNA in situ.<br />

Chromosoma 98, 259–265.<br />

2. Koch, J., Mogensen, J., Pedersen, S., Fischer, H., Hindkjder, J., Kölvraa, S., et al. (1992)<br />

Fast one-step procedure for the detection of nucleic acids in situ by primer-induced<br />

sequence-specific labeling with fluorescein-12-dUTP. Cytogenet. Cell. Genet. 60, 1–3.<br />

3. Gosden, J., Hanratty, D., Starling, J., Fantes, J., Mitchell, A., and Porteous, D. (1991)<br />

Oligonucleotide-primed in situ DNA synthesis (PRINS): A method for chromosome<br />

mapping, banding, and investigation of sequence organization. Cytogenet. Cell. Genet.<br />

57, 100–104.<br />

4. Gosden, J. and Lawson, D. (1994) Rapid chromosome identification by oligonucleotideprimed<br />

in situ DNA synthesis (PRINS). Hum. Mol. Genet. 3, 931–936.<br />

5. Speel, E. J. M., Lawson, D., Hopman, A. H. N., and Gosden, J. (1995) Multi-PRINS:<br />

multiple sequential oligonucleotide primed in situ DNA synthesis reactions label specific<br />

chromosomes and produce bands. Hum. Genet. 95, 29–33.<br />

6. Speel, E. J. M., Lawson, D., Ramaekers, F. C. S., Gosden, J. R., and Hopman, A. H. N.<br />

(1996) Rapid brightfield detection of oligonucleotide primed in situ (PRINS)-labeled DNA<br />

in chromosome preparations and frozen tissue sections. <strong>Bio</strong>Techniques 20, 226–234.<br />

7. Gosden, J. R., ed. (1997) PRINS and In Situ PCR Protocols. Methods in Molecular <strong>Bio</strong>logy,<br />

Vol. 71, Humana Press, Totowa, NJ.<br />

8. Wilkens, L., Tchinda, J., Komminoth, P., and Werner, M. (1997) Single- and double-color<br />

oligonucleotide primed in situ labeling (PRINS): Applications in pathology. Histochem.<br />

Cell <strong>Bio</strong>l. 108, 439– 446.<br />

9. Hindkjder, J., Koch, J., Terkelsen, C., Brandt, C. A., Kölvraa, S., and Bolund, L. (1994)<br />

Fast, sensitive multicolor detection of nucleic acids in situ by primed in situ labeling<br />

(PRINS). Cytogenet. Cell. Genet. 66, 152–154.<br />

10. Speel, E. J. M. (1999) Detection and amplification systems for sensitive, multiple-target<br />

DNA and RNA in situ hybridization: Looking inside cells with a spectrum of colors.<br />

Histochem. Cell <strong>Bio</strong>l. 112, 89–113.


PRINS and Immunocytochemistry 463<br />

11. Kadandale, J. S., Wachtel, S. S., Tunca, Y., Sid Wilroy Jr., R., Martens, P. R., and Tharapel,<br />

A. T. (2000) Localization of SRY by primed in situ labeling in XX and XY sex reversal.<br />

Am. J. Med. Genet. 95, 71–74.<br />

12. Speel, E. J. M., Ramaekers, F. C. S., and Hopman, A. H. N. (1995) Cytochemical detection<br />

systems for in situ hybridization, and the combination with immunocytochemistry. Who is<br />

still afraid of red, green and blue? Histochem. J. 27, 833–858.<br />

13. Speel, E. J. M., Herbergs, J., Ramaekers, F. C. S., and Hopman, A. H. N. (1994) Combined<br />

immunocytochemistry and fluorescence in situ hybridization for simultaneous tricolor<br />

detection of cell cycle, genomic, and phenotypic parameters of tumor cells. J. Histochem.<br />

Cytochem. 42, 961–966.<br />

14. Herbergs, J., Speel, E. J. M., Ramaekers, F. C. S., De Bruïne, A. P., Arends, J. W.,<br />

and Hopman, A. H. N. (1996) Combination of lamin immunocytochemistry and in situ<br />

hybridization for the analysis of chromosome copy numbers in tumor cell areas with high<br />

nuclear density. Cytometry 23, 1–7.<br />

15. Speel, E. J. M., Lawson, D., Ramaekers, F. C. S., Gosden, J. R., and Hopman, A. H.<br />

N. (1997) Combined immunocytochemistry and PRINS DNA synthesis for simultaneous<br />

detection of phenotypic and genomic parameters in cells, in PRINS and in situ PCR<br />

protocols, Methods in Molecular <strong>Bio</strong>logy, Vol. 71 (Gosden, J. R., ed.), Humana Press,<br />

Totowa, NJ, pp. 53–60.<br />

16. Roy, L., Coullin, P., Vitrat, N., Hellio, R., Debili, N., Weinstein, J., et al. (2001) Asymmetrical<br />

segregation of chromosomes with a normal metaphase/anaphase checkpoint in<br />

polyploid megakaryocytes. Blood 97, 2238–2247.<br />

17. Coullin, P., Roy, L., Pellestor, F., Candelier, J.-J., Bed’’hom, B., Guillier-Gencik, Z., et al.<br />

(2002) PRINS, the other in situ DNA labeling method that is useful in cellular biology.<br />

Am. J. Med. Genet. 107, 127–135.<br />

18. Jeppesen, P. (1994) Immunofluorescence techniques applied to mitotic chromosome<br />

preparations, in Chromosome Analysis Protocols, Methods in Molecular <strong>Bio</strong>logy (Gosden,<br />

J. R., ed.), Humana Press, Totowa, NJ, pp. 253–286.<br />

19. Mitchell, A. R. (1997) Chromosomal PRINS DNA labeling combined with indirect<br />

immunocytochemistry, in PRINS and in situ PCR protocols, Methods in Molecular <strong>Bio</strong>logy<br />

(Gosden, J. R., ed.), Humana Press, Totowa, NJ, pp. 61–70.<br />

20. Glaser, T., Housman, D., Lewis, W. H., Gerhard, D., and Jones, C. (1989) A fine-structure<br />

deletion map of chromosome 11p: analysis of J1 series hybrids. Somat. Cell Mol. Genet.<br />

15, 477–501.<br />

21. Boerman, O. C., Mijnheere, E. P., Broers, J. L. V., Vooijs, G. P., and Ramaekers, F. C. S.<br />

(1991) <strong>Bio</strong>distribution of a monoclonal antibody (RNL-1) against the neural cell adhesion<br />

molecule (NCAM) in athymic mice bearing human small-cell lung-cancer xenografts. Int.<br />

J. Cancer 48, 457– 462.<br />

22. Dorin, J. R., Inglis, J. D., and Porteous, D. J. (1989) Selection for precise chromosomal<br />

targeting of a dominant marker by homologous recombination. Science 243, 1357–1360.<br />

23. Carney, D. N., Gazdar, A. F., Bepler, G., Guccion, J. G., Marangos, P. J., Moody, T. W.,<br />

et al. (1985) Establishment and identification of small cell lung cancer cell lines having<br />

classic and variant features. Cancer Res. 45, 2913–2923.<br />

24. Koch, J., Hindkjder, J., Kölvraa, S., and Bolund, L. (1995) Construction of a panel of<br />

chromosome-specific oligonucleotide probes (PRINS-primers) useful for the identification<br />

of individual human chromosomes in situ. Cytogenet. Cell. Genet. 71, 142–147.<br />

25. Pellestor, F., Girardet, A., Lefort, G., Andréo, B., and Charlieu, J.-P. (1995) Selection<br />

of chromosome specific primers and their use in simple and double PRINS technique<br />

for rapid in situ identification of human chromosomes. Cytogenet. Cell. Genet. 70,<br />

138–142.


464 Speel, Ramaekers, and Hopman<br />

26. Krejci, K. and Koch, J. (1999) An in situ study of variant telomeric repeats in human<br />

chromosomes. Genomics 58, 202–206.<br />

27. Ramael, M., Van Steelandt, H., Stuyven, G., Van Steenkiste, M., and Degroote, J. (1999)<br />

Detection of human papilloma virus (HPV) genomes by the primed in situ labelling<br />

technique. Pathol. Res. Pract. 195, 801–807.<br />

28. Bobrow, M. N., Harris, T. D., Shaughnessy, K. J., and Litt, G. J. (1989) Catalyzed<br />

reporter deposition, a novel method of signal amplification. Application to immunoassays.<br />

J. Immunol. Methods 125, 279–285.<br />

29. Speel, E. J. M., Hopman, A. H. N., and Komminoth, P. (1999) Amplification methods to<br />

increase the sensitivity of in situ hybridization. Play card(s). J. Histochem. Cytochem.<br />

47, 281–288.<br />

30. Speel, E. J. M., Hopman, A. H. N., and Komminoth, P. (2000) Signal amplification for DNA<br />

and mRNA in situ hybridization, in In Situ Hybridization Protocols, Methods in Molecular<br />

<strong>Bio</strong>logy (Darby, I. A., ed.), Humana Press, Totowa, NJ, pp. 195–216.


Cloning and Mutagenesis 467<br />

64<br />

Cloning and Mutagenesis<br />

A Technical Overview<br />

Helen Pearson and David Stirling<br />

1. Introduction<br />

Strategies for cloning polymerase chain reaction (PCR) products and performing in<br />

vitro site-directed mutagenesis are legion, and the following chapters outline five robust<br />

and reliable protocols. Before embarking on such a strategy, however, it is worth considering<br />

if it is entirely necessary. Even high-fidelity, proofreading Taq polymerases carry the<br />

risk of misincorporated bases being included, especially late in the PCR, when dNTP<br />

concentrations may become limiting. Thus, if the product is cloned, there is a chance<br />

that the clone selected contains a misincorporated base. If the purpose of the exercise is<br />

simply to determine the sequence of the original template, it may be more appropriate<br />

to sequence the PCR product directly and so avoid such cloning bias. If the object is to<br />

produce a clone that can be further used in expression studies for instance, it is imperative<br />

that all cloned material is sequenced to verify its integrity, prior to expression.<br />

2. Cloning<br />

2.1. T Overhangs<br />

Taq polymerases lacking 3′ to 5′ exonuclease activity tends to add nontemplatedirected<br />

nucleotides to the ends of double-stranded DNA fragments. If blunt end<br />

cloning is to be used, these overhangs need to be “polished,” and Chapter 65 by<br />

Kai Wang provides an optimized method using T4 DNA polymerase. Because the<br />

predominant nucleotide added in this nontemplate-directed manner is adenosine, many<br />

successful protocols have used vectors with a thymidine overhang to direct the cloning.<br />

This allows rapid cloning into a shuttle vector, from which the fragment can then be<br />

restricted and cloned further. Alternatively, Horton and colleagues present an elegant<br />

protocol (Chapter 66) using T-linkers to enable cloning into specific sites.<br />

2.2. Restriction Site Primers<br />

Neither blunt-end cloning nor the use of T-overhang vectors allows directional<br />

cloning. If the region of template DNA to be amplified contains suitable restriction<br />

sites, the product can simple be digested and cloned in the same way as any other DNA<br />

fragment. If this is not feasible, it is possible to introduce restriction site sequences<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

467


468 Pearson and Stirling<br />

into PCR products by having these sequences incorporated into the 5′ end of the PCR<br />

primer(s). The short restriction site sequence on the 5′ end of the PCR primer will not<br />

hybridize, but as long as the 3′ hybridizing region is long enough (i.e., its T m is high<br />

enough; ~20 mer), the primer will specifically bind to the appropriate site. The PCR<br />

product will thus have an additional DNA sequence at the 5′ end that will contain the<br />

endonuclease restriction site. A similar or different restriction site sequence can be<br />

added via the other PCR primer. If the other primer has a different restriction sequence,<br />

then the PCR fragment can be inserted in a directional-dependent manner in a host<br />

plasmid. There are a number of potential problems with this method that should be<br />

considered. There is no easy way to prevent internal sites containing similar restriction<br />

sequences from being cut when the end of the PCR product are cut. Care should<br />

therefore be taken to use restriction sites that are not present in the fragment to be<br />

amplified. Restriction sequences are inverse repeat sequences, thus the potential exists<br />

for primer dimer association and resultant non-productive annealing. Finally, when<br />

restriction sites are located very close to the end of an amplified fragment, the efficiency<br />

of cleavage of those sites can be markedly impaired. It may therefore be necessary<br />

to include not just the restriction site but an additional 5 to 10 bases to avoid this<br />

problem.<br />

3. Mutagenesis<br />

In vitro site-directed mutagenesis is an invaluable technique for studying protein<br />

structure–function relationships, gene expression, and vector modification. Several<br />

methods have appeared in the literature, but many of these methods require singlestranded<br />

DNA as the template. The reason for this, historically, has been the need<br />

for separating the complementary strands to prevent reannealing. Use of PCR in<br />

site-directed mutagenesis accomplishes strand separation by using a denaturing step<br />

to separate the complementing strands and allowing efficient polymerization of the<br />

PCR primers. PCR site-directed methods thus allow site-specific mutations to be<br />

incorporated in virtually any double-stranded plasmid, eliminating the need for M13-<br />

based vectors or single-stranded rescue.<br />

Three divergent strategies for mutagenesis are outlined in the following chapters;<br />

however, several points applicable to all three should be here. First, it is often desirable<br />

to reduce the number of cycles during PCR when performing PCR-based site-directed<br />

mutagenesis to prevent clonal expansion of any (undesired) second-site mutations.<br />

Limited cycling, which would result in reduced product yield, can be offset by<br />

increasing the starting template concentration. Second, a selection must be used to<br />

reduce the number of parental molecules coming through the reaction. This is of<br />

particular importance when the parental molecules are used in high concentrations.<br />

Third, because of the tendency of some thermostable polymerases to add nontemplatedirected<br />

nucleotides to the ends of double-stranded DNA fragments, it is often necessary<br />

to incorporate an end-polishing step into the procedure prior to end-to-end ligation of<br />

the PCR-generated product containing the incorporated mutations in one or both PCR<br />

primers (see Subheading 2.1.).<br />

Finally, even if the presence of the desired mutation is confirmed by restriction<br />

digest or sequencing, it is essential to sequence the entire region of manipulated<br />

DNA to ensure that there has been no undesirable mutation introduced by the PCR<br />

processes.


T4 DNA Polymerase 469<br />

65<br />

Using T4 DNA Polymerase<br />

to Generate Clonable PCR Products<br />

Kai Wang<br />

1. Introduction<br />

Polymerase chain reaction (PCR) mediated through Taq DNA polymerase has<br />

become a simple and routine method for cloning, sequencing, and analyzing genetic<br />

<strong>info</strong>rmation from very small amounts of materials (1). Taq DNA polymerase, like some<br />

other DNA polymerases, lacks 3′ to 5′ exonuclease activity and will add nontemplatedirected<br />

nucleotides to the ends of double-stranded DNA fragments. Because of the<br />

strong preference of the Taq polymerase for dATP, the nucleotide added is almost<br />

exclusively an adenosine (2). This results in generating “ragged” unclonable amplification<br />

products (2,3). Restriction endonuclease sites are often incorporated into the<br />

amplification primers so that clonable PCR products can be generated by restriction<br />

enzyme cleavage (4). However, the possible secondary sites located within amplified<br />

products often complicate the cloning and interpretation of PCR results. A cloning<br />

system exploiting the template-independent terminal transferase activity of Taq<br />

polymerase has been reported (5–7). However, a special vector with thymidine (T)<br />

overhanging ends has to be used in the process.<br />

T4 DNA polymerase has very strong exonuclease and polymerase activities in a<br />

broad range of reaction conditions (8). By adapting its strong enzymatic activities,<br />

a simple and efficient method to generate clonable PCR fragments with T4 DNA<br />

polymerase has been developed (9). The T4 DNA polymerase not only repairs the ends<br />

of the PCR products, but it also removes the remaining primers in the reaction with its<br />

strong single-stranded exonuclease activity. Therefore, this method usually does not<br />

require multiple sample handling, buffer changes, or gel purification steps. Instead, a<br />

simple alcohol precipitation step is used to purify the PCR products.<br />

The blunt-end cloning protocol can be modified for sticky-end cloning. Even<br />

though this may increase cloning efficiency to a certain extent, a purification step,<br />

to remove excess deoxynucleotides from PCRs, has to be added before adding T4<br />

DNA polymerase.<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

469


470 Wang<br />

2. Materials<br />

2.1. PCR<br />

1. DNA template containing the sequence of interest.<br />

2. Oligonucleotide primers.<br />

3. Taq DNA polymerase.<br />

4. 10× PCR and enzymatic repair buffer: 500 mM Tris-HCl, pH 9.0, 25 mM MgCl 2 , 500 mM<br />

NaCl, and 5 mM DTT. Commercial 10× PCR buffer also works well.<br />

5. 1.5 mM 10× deoxynucleotides (dNTP) solution. Concentrated stock solution (100 mM ) can<br />

be obtained from Amersham <strong>Bio</strong>sciences (Piscataway, NJ) or Boehringer Mannheim<br />

(Indianapolis, IN).<br />

6. Gel electrophoresis and PCR equipment.<br />

2.2. End Repair and Blunt-End Cloning<br />

1. Enzymes (T4 DNA polymerase, T4 polynucleotide kinase, and T4 DNA ligase). Enzymes<br />

can be purchased from Boehringer Mannheim, Invitrogen (Carlsbad, CA) or any other<br />

provider.<br />

2. 1 mM of ATP solution. Concentrated stock solution (100 mM) can be obtained from<br />

Boehringer Mannheim.<br />

3. Isopropyl alcohol.<br />

4. Vector (blunt end and dephosphorylated).<br />

5. 10× Ligase buffer: 660 mM Tris-HCl, pH 7.6, 66 mM MgCl 2 , 10 mM ATP, 1 mM spermidine,<br />

and 10 mM DTT. Commercially available 10× ligase buffers also works well.<br />

6. TE buffer: 10 mM Tris-HCl, pH 7.5, 1 mM EDTA.<br />

7. 5 M NaCl.<br />

2.3. End Repair and Sticky-End Cloning<br />

1. Sephacryl S-400 spin column. A commercial spin column (MicroSpin S-400HR) can be<br />

obtained from Pharmacia.<br />

2. Enzymes (T4 DNA polymerase, T4 polynucleotide kinase, and T4 DNA ligase). Enzymes<br />

can be purchased from Boehringer Mannheim, Invitrogen, or any other provider.<br />

3. ATP and dNTPs.<br />

4. Isopropyl alcohol.<br />

5. Vector (digested and dephosphorylated).<br />

6. 10× Ligase buffer: 660 mM Tris-HCl, pH 7.6, 66 mM MgCl 2 , 10 mM ATP, 1 mM spermidine,<br />

and 10 mM DTT. Commercially available 10× ligase buffers also works well.<br />

7. 0.5 M EDTA, pH 8.0.<br />

8. 5 M NaCl.<br />

3. Method<br />

3.1. Primer Design<br />

For blunt-end cloning, no special primer is needed. However, secondary structure<br />

and stretch of homopolymer should be avoided. For sticky-end cloning, depending on<br />

restriction site selected, specific sequences should be included so that compatible ends<br />

can be generated after T4 DNA polymerase treatment (see Subheading 3.4.).<br />

3.2. PCR<br />

1. Prepare the following in a PCR tube: 5 µL of 10× PCR buffer; 0.25 µg of genomic DNA;<br />

1 µM of each primer; 0.15 mM of each deoxynucleotide (5 µL of 1.5 mM stock dNTP<br />

solution), 1 U Taq polymerase, and deionized H 2 O to a final volume of 50 µL.


T4 DNA Polymerase 471<br />

2. Amplification conditions largely depend on the specific applications. However, a general<br />

cycling profile can be used in most experiments: 94°C for 7 min (initial denaturation); 94°C<br />

for 30 s (amplification), 55°C for 45 s, 72°C for 90 s; and 72°C for 10 min (extension).<br />

3. Examine the PCR amplification results with agarose gel electrophoresis (see Note 1).<br />

3.3. End Repair and Blunt-End Cloning<br />

1. Add the following to PCR tubes directly to repair the PCR products (see Note 2): 1 U<br />

of T4 DNA polymerase; 1 UL of 4 mM dNTP solution (optional; see Note 3); 5 U of T4<br />

polynucleotide kinase (see Note 4); and 1 µL of 1 MM ATP.<br />

2. Incubate the reaction tubes at 25°C (room temperature) for 20 min (see Notes 5 and 6).<br />

Stop the reactions by adding 3 mL of 0.5 M EDTA, pH 8.0.<br />

3. Incubate the reaction tubes at 70°C for 10 min to inactive the enzymes.<br />

4. Precipitate the PCR products by adding 5 µL of 5 M NaCl and 60 µL of isopropyl alcohol<br />

(see Notes 7, ref. 8).<br />

5. Resuspend the DNA fragments in 20 µL of TE or water.<br />

6. Take 2 µL of DNA (containing approx 20–50 ng of PCR product) and mix with ligase<br />

and vector for ligation (see Note 8): 1 µL of 10× ligase buffer; 1 µL of T4 ligase;<br />

2 µL of repaired DNA; dephosphorylated vector (60–150 ng); and add deionized H 2 O<br />

to a final volume of 10 µL.<br />

7. Incubate at 16°C overnight.<br />

8. Dilute the ligation reaction fivefold in TE buffer. Use 2 µL of the diluted ligation reaction<br />

for transformation (see Note 9).<br />

3.4. End Repair and Sticky-End Cloning<br />

An EcoRI site is used as an example in the following protocol Fig. 1). However,<br />

depending on the desired cloning site, a different combination of dNTP should be<br />

added in the “repair” reaction (see Subheading 3.4., step 2).<br />

1. Spin through the PCR mixture (40 µL) in a pre-equilibrated Sephacryl S-400HR spin<br />

column (see Note 10) to remove excess dNTP.<br />

2. Add the following to the column-purified PCR fragments to generate sticky ends: 5 µL<br />

of 10× PCR buffer; 1 U of T4 DNA polymerase; 5 U of T4 polynucleotide kinase; 1 µL<br />

of 4 MM dCTP and dGTP; 1 µL of 1 MM ATP; and add deionized H 2 O to a final volume<br />

of 50 µL.<br />

3. Incubate the reaction tubes at 25°C (room temperature) for 20 min. Stop the reaction by<br />

adding 3 µL of 0.5 M EDTA, pH 8.0.<br />

4. Heat inactivate the enzymes by placing the reaction tubes at 70°C for 10 min.<br />

5. Precipitate the PCR products by adding 5 µL of 5 M NaCl and 60 µL of isopropyl<br />

alcohol (8).<br />

6. Resuspend the DNA fragments in 20 µL of TE or water.<br />

7. Take 2 µL of DNA (containing approx 20–50 ng of DNA) and mix with ligase and EcoRI<br />

digested, dephosphorylated vector for cloning: 1 µL of 10× ligase buffer; 1 µL of T4<br />

ligase; 2 µL of repaired DNA; dephosphorylated vector (60–150 ng); and add deionized<br />

H 2 O to a final volume of 10 µL<br />

8. Incubate at 16°C overnight.<br />

9. Dilute the ligation reaction fivefold in TE buffer. Use 2 µL of the diluted ligation reaction<br />

for transformation (see Note 9).<br />

4. Notes<br />

1. In case of multiple PCR products from a single reaction, the specific products should<br />

be purified by gel electrophoresis based on size after repair reaction. Several different


472 Wang<br />

Fig. 1. A brief outline of the strategy used to generate sticky-end PCR products with T4<br />

DNA polymerase.<br />

methods can be used to purify DNA fragments from agarose gel, such as phenol extraction<br />

from low-melting gel (8), the “glassmilk” method, or simple low-speed centrifugation<br />

(10). The phenol extraction method has been found to be less expensive and able to recover<br />

a sufficient amount of clean DNA for cloning.<br />

2. This protocol uses one buffer for all the enzymes that include Taq polymerase in PCR,<br />

T4 polymerase, and T4 polynucleotide kinase in end-repair reaction. Therefore, a slightly<br />

higher concentration of reagents and enzymes can be added in the reaction.<br />

3. T4 DNA polymerase can be added directly into the PCR tube without providing additional<br />

nucleotides. However, T4 DNA polymerase balances its exonuclease and polymerase<br />

activities based on the concentration of available deoxynucleotides. Depending on the<br />

length of amplification products, number of amplification cycles, and nucleotide sequence<br />

composition of amplified region, the remaining nucleotide concentration after PCR<br />

amplification may be different from experiment to experiment. To avoid unnecessary<br />

confusion, supplemental nucleotides are routinely added for end-repair reaction.<br />

4. T4 polynucleotide kinase is not needed when vector used has not been treated with<br />

phosphatase previously. However, dephosphorylated vector should be used to increase<br />

the cloning efficiency.


T4 DNA Polymerase 473<br />

5. Room temperature (25°C) was chosen for the reaction since T4 DNA polymerase has<br />

excessive exonuclease activity at 37°C.<br />

6. The T4 polynucleotide kinase works well at room temperature as opposed to the higher<br />

reaction temperature (37°C) regularly used (8).<br />

7. Although the PCR products purified directly by alcohol precipitation after end repairing<br />

are sufficient for routine cloning, passing the repaired PCR product mixtures through a<br />

gel filtration column prior to the alcohol precipitation can greatly enhance the cloning<br />

efficiency.<br />

8. In the ligation reaction, we routinely used 11 molar ratio between vector (dephosphorylated)<br />

and insert. Generally more than 200 recombinant clones can be obtained with 0.4 µL<br />

of the ligation reaction. Therefore, a single ligation reaction for each PCR product is<br />

sufficient for most applications.<br />

9. Because of its reliability and high transformation efficiency, commercial CaCl 2 -treated<br />

competent cells are used for the transformation step. The bacteria strain we routinely used<br />

is DH10B (BRL; MAX efficiency DH10B competent cells). However, various competent<br />

cells can be purchased from companies, such as Stratagene and Invitrogen.<br />

10. Prepacked Sephacryl S-400HR spin columns (MicroSpin S-400HR) can be purchased<br />

from Pharmacia. Alternatively, the spin columns can be prepared from bulk gel filtration<br />

matrix (Sephacryl S-400HR, Pharmacia) as described in other protocol books (8). The<br />

filtration medium (prepacked or bulk filtration matrix) contains 20% alcohol; therefore,<br />

the spin columns should be washed and equilibrated with TE.<br />

References<br />

1. Saiki, R. K., Scharf, S., Faloona, F., Mullis, K. B., Horn, G. T., Erlich, H. A., et al. (1985)<br />

Primer-directed enzymatic amplification of DNA with a thermostable DNA polymerase.<br />

Science 230, 1350–1354.<br />

2. Clark, J. M. (1988) Novel non-templated nucleotide addition reactions catalyzed by<br />

procaryotic and eucaryotic DNA polymerases. Nucleic Acids Res. 16, 9677–9686.<br />

3. Scharf, S. (1990) PCR Protocols: A Guide to Methods and Applications, Academic Press,<br />

San Diego, CA.<br />

4. Jung, V., Pestka, S. B., and Pestka, S. (1990) Efficient cloning of PCR generated DNA<br />

containing terminal restriction endonuclease recognition sites. Nucleic Acids Res. 18,<br />

6156.<br />

5. Mead, D. A., Pey, N. K., Herrnstadt, C., Marcil, R. A., and Smith, L. M. (1991) A universal<br />

method for the direct cloning of PCR amplified nucleic acid. <strong>Bio</strong>/Technology 9, 657.<br />

6. Kovalic, D., Kwak, J., and Weisblum, B. (1991) General method for direct cloning of DNA<br />

fragments generated by the polymerase chain reaction. Nucleic Acids Res. 19, 4560.<br />

7. Marchuk, D., Drumm, M., Saulino, A., and Collins, F. (1991) Construction of T-vectors,<br />

a rapid and general system for direct cloning of unmodified PCR products. Nucleic Acids<br />

Res. 19, 1154.<br />

8. Sambrook, J., Fritsch, E. F., and Maniatis, T. (1989) Molecular Cloning: A Laboratory<br />

Manual, 2nd ed., Cold Spring Harbor Laboratory, Cold Spring Harbor, NY.<br />

9. Wang, K., Koop, B. F., and Hood, L. (1994) A simple method using T4 DNA polymerase to<br />

clone polymerase chain reaction products. <strong>Bio</strong>techniques 17, 236–238.<br />

10. Heery, D. M., Gannon, F., and Powell, R. (1990) A simple method for subcloning DNA<br />

fragments from gel slices. Trends Genet. 6, 173.


474 Wang


T-Linker Strategy 475<br />

66<br />

A T-Linker Strategy for Modification<br />

and Directional Cloning of PCR Products<br />

Robert M. Horton, Raghavanpillai Raju, and Bianca M. Conti-Fine<br />

1. Introduction<br />

The propensity of Taq polymerase to add 3′-A overhangs (1,2) to polymerase chain<br />

reaction (PCR)-amplified DNA has made possible a simple method for cloning PCR<br />

products into a T-vector (Invitrogen, San Diego, CA) (3–5). Here, we present a related<br />

strategy that uses T-linkers to add sequences, such as restriction sites, to the ends<br />

of PCR products (see Note 1). A single-base T overhang at the end of a synthetic<br />

double-stranded oligonucleotide linker allows ligation of the linker to the unpolished<br />

ends of a PCR product. This avoids the expense of adding the “extra” sequences to<br />

sequence-specific primers.<br />

1.1. Examples<br />

Two T-linker designs are presented here. In each case, the T-linker is a doublestranded<br />

synthetic oligonucleotide composed of complementary oligos (either TL-A<br />

and TL-B for the NdeT linker or HisTL-A and HisTL-B for the HisT-linker) with a<br />

single 3′ overhanging t at one end.<br />

1.1.1. NdeTL<br />

The basic principles involved in using a T-linker are shown using the Nde-T-linker<br />

in Fig. 1. This T-linker contains complete EcoRI and NotI sites, and a third site (NdeI)<br />

is partially present, except for its final g; the overhanging t is part of this site. The 5′<br />

end of TL-B is phosphorylated (indicated by an asterisk) so that it can be ligated. The<br />

other end of the linker contains a sticky HindIII-compatible end, which was not used<br />

in the approach described here. However, because this 5′ overhang is filled in by the<br />

polymerase during PCR, these extra bases serve as a “clamp” or spacer, which permits<br />

the EcoRI site to be cut.<br />

The overhanging t of the linker matches the a added by the polymerase and directs<br />

the ligation of TL-B, allowing reamplification of the sequence using TL-A as a primer.<br />

The resulting molecule contains the original PCR-amplified sequence flanked by<br />

inverted repeats of the T-linker.<br />

If one of the sequence-specific primers has a g at its 5′ end, an NdeI site will be<br />

formed. This site is “split” between the T-linker and the PCR product to be cloned. In<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

475


476 Horton, Raju, and Conti-Fine<br />

Fig. 1. Cloning of PCR products with a T-linker. (A) Oligonucleotide design. (B) Addition<br />

of the T-linker to a PCR product. The parenthetical g and NdeI indicate that if a g is the first base<br />

in the PCR product, the NdeI site will be completed. (C) Sequence from a pBluescript KS II+<br />

plasmid containing a T-linker-reamplified insert cloned into its EcoRI site. This sequence was<br />

obtained using an fmol cycle sequencing kit (Promega, Madison, WI). The base representing<br />

the overhanging t of the T-linker is indicated by the solid triangle to the right of the sequence.<br />

Since the PCR product begins with a g, a complete NdeI site is formed.<br />

this way, very minor restraints in the PCR primer sequences (having one start with g,<br />

but not the other) can be used to complete the site at only one end of the reamplified<br />

product. One simple application of this idea is directional cloning (see Notes 2 and 3).<br />

The use of this T-linker is illustrated in Fig. 1. A portion of a T-cell receptor V region<br />

was amplified from Jurkat tumor line cDNA using primers Vb8-1cpe5 and hpVbe3.<br />

The ligated products were reamplfied with TL-A and cloned into the EcoRI site of<br />

pBluescript KS II + . The sequence of the resulting product is shown in Fig. 1C. At<br />

the bottom of the gel is sequence from the polylinker of the vector to the EcoRI site,<br />

followed by the T-linker. The overhanging t is marked. The g completing the NdeI site<br />

is the first base of Vb81cp.5.<br />

1.1.2. HisTL<br />

Our second example adds another level of sophistication, as shown in Fig. 2. Here<br />

an additional ability of the T-linker is brought into use. A new and useful sequence<br />

is added in addition to the restriction sites, namely a “histidine tag” sequence, which<br />

will be used for affinity purification of the expressed recombinant α3 protein on a Ni 2+<br />

column (Qiagen, Chatsworth, CA). Note that the HindIII site is created only at the<br />

3′ end of the product in this case, because only the 3′ primer begins with t. The<br />

histidine tag portion is removed from the 3′ end by cutting with HindIII. No special<br />

sequences have been added to the sequence-specific primers hacpe5 and hacpe3, but<br />

their positions have been chosen to take advantage of the design of the HisT-linkers.<br />

The initial PCR-amplified sequence begins with a full codon to put the sequence in


T-Linker Strategy 477<br />

Fig. 2. Using a T-linker to add a “histidine tag” to one end of a PCR product. (A) Oligonucleotide<br />

design. The NdeI site allows the product to be cloned into an appropriate expression<br />

vector, in which translation begins at the ATG codon included in this site. The amino acids that<br />

make up the histidine tag are shown, and the reading frame is indicated by vertical bars between<br />

codons. The NotI site in HisTL-A is in parentheses because it would not be expected to cut so<br />

close to the end of a DNA molecule (although it should be possible to use it if the primer were<br />

phosphorylated and the products ligated to form concatamers). Its main purpose is to serve as<br />

a clamp to allow efficient cutting at the NdeI site. (B) Sequence at the junction between the<br />

HisT-linker and the 5′ end of the PCR-amplified sequence. The 5′ primer begins at the first base<br />

in a codon, to put the sequence in frame with the histidine tag. (C) Sequence at the 3′ junction.<br />

The 3′ primer begins with ta; the complementary ta in the other strand is converted to a stop<br />

codon “taa” when the extra a is added by Taq polymerase. Because the 5′ primer begins with c,<br />

the HindIII site is completed only at the 3′ end of the molecule. This allows directional cloning<br />

of the product, and removal of the T-linker sequences from the 3′ end.


478 Horton, Raju, and Conti-Fine<br />

the proper reading frame with the His tag. It ends with “ta”; the “a” added by the<br />

polymerase finishes the termination codon (taa) and is included in the HindIII site.<br />

2. Materials<br />

1. cDNA template: This was reverse transcribed from total RNA using Superscript RNase<br />

H-reverse transcriptase (Life Sciences) and an oligo (dT) primer using manufacturers<br />

instructions.<br />

2. Reagents for PCR: 10× buffer: 500 mM KCl, 100 mM Tris-HCl, pH 8.3; red sucrose is<br />

a PCR-compatible gel-loading dye consisting of ~1 mM cresol red in 60% sucrose (6);<br />

Taq DNA polymerase, 10 mM dNTP mix, 10 mM MgCl 2 solution (Perkin–Elmer Cetus,<br />

Norwalk, CT).<br />

3. Reagents for agarose gel electrophoresis.<br />

4. GeneClean (<strong>Bio</strong> 101, La Jolla, CA).<br />

5. Ligation reagents: T4 DNA ligase (Stratagene, La Jolla, CA) 8 U/µL, ligase buffer supplied<br />

by manufacturer, 10 mM ATP (Pharmacia, Piscataway, NJ), and PEG 8000 (Aldrich,<br />

Milwaukee, WI). Recently, we have used T4 ligase buffer from Life Technologies (Gibco-<br />

BRL, Gaithersburg, MD), which already contains ATP and PEG.<br />

6. T-linker oligonucleotides:<br />

a. Nde-T-linkers: TL-A: (23-mer) 5′-agcttgaattcgcggccgcatat-3′; TL-B: (18-mer)*<br />

5′-tatgcggccgcgaattca-3′;<br />

b. His-T-linkers: HisTL-A: (55-mer) 5′-gcggccgcatatgggatcctcacatcatcatcaccatcactcgagtggccaagct-3′;<br />

HisTL-B: (21-mer) *5′-gcttggccactcgagtgatgg-3′<br />

The 5′ end of each “B” oligonucleotide is phosphorylated (represented by the *) so<br />

that it can be ligated to the (nonphosphorylated) end of the PCR product. The 5′ end<br />

of the “A” oligonucleotide is not phosphorylated. The “B” oligo only needs to be long<br />

enough to bind the “A” oligo during the ligation and in the annealing steps of the early<br />

rounds of reamplification. A 5′ overhang on oligo “A” is not a problem, because this is<br />

filled in by the polymerase during reamplification.<br />

c. Dissolve primers at a stock concentration of 10 µM, which is generally considered<br />

20× for PCR. After mixing T-linker primers, they are at a final concentration of approx<br />

5 µM for ligation reactions.<br />

7. Ampligrease (see ref. 7): plain petroleum jelly, Vaseline brand, or generic is suitable.<br />

Apply quality control checks as described (7).<br />

8. Reagents for bacterial culture, including competent Escherichia coli, appropriate antibiotic,<br />

LB agar.<br />

9. 95% ethanol containing 2% (w/v) potassium acetate.<br />

10. 75% ethanol.<br />

3. Methods<br />

3.1. PCR<br />

The HisT-linker was used to clone the coding region of the α3 subunit of the<br />

nicotinic acetylcholine receptor from human bronchial epithelium (manuscript in<br />

preparation; see Note 2) into the E. coli expression vector pT7-7 (8). The following<br />

sequence-specific primers were used:<br />

Vb8-1cpe5:<br />

hpVbe3:<br />

ha3cpe5:<br />

ha3cpe3:<br />

5′-ggagttatccagtcacc-3′<br />

5′-gggaattcgtcgactgctggcrcagarrta-3′<br />

5′-ccagtggccagggcctcaga-3′<br />

5′-tatgcatcttccctggccatca-3′


T-Linker Strategy 479<br />

1. Perform PCRs under fairly standard conditions. For example, for cloning the α3 subunit<br />

of the nicotinic receptor, the conditions for amplification were as follows: 3 µL of 10×<br />

PCR buffer, 4.5 µL of 10 mM MgCl 2 , 3 µL of dNTPs (2 mM each), 6 µL of red sucrose,<br />

16.5 µL of H 2 O, 0.75 µL of ha3cpe5 primer (10 µM), 0.75 µL of ha3cpe3 primer (10 µM ),<br />

1 µL of cDNA template, and 0.25 µL of Taq polymerase.<br />

2. Perform PCR using 40 cycles at 94, 52, and 72°C, each for 0.5 min, in a programmable<br />

circulating air oven (ProOven, Integrated Separation Systems).<br />

3. Products were purified by agarose gel electrophoresis and GeneClean (<strong>Bio</strong> 101; see<br />

Chapter 18).<br />

3.2. Ligation (see Note 4)<br />

1. Assemble ligation reactions as follows: 2 µL of purified PCR product, 7.5 µL of H 2 O,<br />

2 µL of 10× ligase buffer, 1 µL of (0.5 mM final) 10 mM ATP (if not in buffer), 2.5%<br />

(5% final) 40% PEG 8000 (optional), 2 µL of T-linkers (5 µM in 500 mM NaCl), and<br />

1 µL of ligase (1 U/µL dilution).<br />

2. Incubate at room temperature for 20 min to overnight.<br />

3.3. Reamplification with T-Linker Primer<br />

1. After ligation of the T-linker, reamplfy the products as follows: 5 µL of 10× PCR buffer, 7.5<br />

µL of 10 mM MgCl 2 , 5 µL of dNTPs, 10 µL of Red sucrose, 16.5 µL of H2O, 5 µL of TL-A<br />

(or HisTL-A) primer (10 µM ), ligation mixture, and 0.35 µL of Taq polymerase.<br />

2. Cover the sample with mineral oil and amplify with 40 two-step cycles of 94°C for<br />

0.5 min and 72°C for 2.5 min. (see Note 5).<br />

3.4. Alternative Method: One-Tube Ligation/Reamplification<br />

The whole T-linker reaction can be set up in one tube to make a “kit-like” product.<br />

This is done using a meltable barrier, as for hot-start PCR (we use AmpliGrease; ref. 7).<br />

The “top mix” contains the ligation reaction, to which the PCR-amplified band is<br />

added. After a suitable incubation to allow ligation, the reaction is heated to melt the<br />

barrier, and the second PCR is begun:<br />

1. Set up a 100-µL PCR mix (see Subheading 3.1.) in a tube. Add only TL-A (or HisTL-A)<br />

as a primer, and no template.<br />

2. Dispense approx 35 µL of petroleum jelly (AmpliGrease) onto the side of the tube with<br />

a syringe.<br />

3. Heat the tube so that the grease melts to cover the bottom mix.<br />

4. Allow to cool so that the grease resolidifies.<br />

5. Add approx 30 µL of mineral oil on top of the grease.<br />

6. Add 4 µL of ligation mix through the oil so the droplet rests above the grease barrier. The<br />

ligation mix is made as a master mix containing the following: 2 µL of T-linkers (5 µM each<br />

in 50 mM NaCl), 2 µL of 10× ligase buffer, 1 µL of 10 mM ATP (if not in buffer), 0.5 µL<br />

of ligase (8 U/µL), and 10.5 µL of H 2 O.<br />

7. Add 1 µL of the PCR product to be cloned into the droplet of top ligation mix.<br />

8. Use the following reaction conditions: 25°C for 1 h (ligation); 94°C for 15 s, 55°C for 15 s,<br />

and 72°C for 45 s for 30 amplification cycles (see Note 6).<br />

3.5. Digestion and Cloning<br />

The reamplified PCR product now has the restriction sites from the T-linker at its<br />

ends, and may be cloned using standard procedures. The protocol we usually use is<br />

as follows:


480 Horton, Raju, and Conti-Fine<br />

1. Run a small portion of the reamplified product on a checking gel to make sure you have<br />

enough of the correctly sized product with which to work. About 5 µL should give a clear<br />

band. With care, you should be able to use this same gel to check your DNA at several<br />

stages of the process. Do not let it dry out!<br />

2. Add ~1 µg of uncut (supercoiled) plasmid vector to the tube containing the reamplified<br />

material. This will both act as a carrier during the purification process and help to make<br />

the effects of the restriction digests more obvious. You will probably not be able to see<br />

the slight size differences resulting from cutting off the ends of the T-linkers, but you<br />

should be able to see the difference as the vector is cut, and goes from being supercoiled<br />

to being linear.<br />

3. Extract the DNA twice with phenol:chloroform and once with chloroform.<br />

4. Precipitate the DNA by adding 2 vol of 95% ethanol containing 2% potassium acetate to<br />

the aqueous supernatant and spin on high speed in a microcentrifuge at 4°C for 30 min.<br />

Discard the supernatant and carefully rinse the pellet with cold 75% ethanol.<br />

5. Resuspend the DNA pellet in distilled water and add the appropriate 10× restriction<br />

enzyme buffer. Add restriction enzymes and incubate until the vector is completely in the<br />

linear form on a checking gel. Directional cloning will require cutting with two enzymes.<br />

Manufacturers generally provide charts that indicate which buffer works reasonably well<br />

with both enzymes. Run a lane with uncut vector on the checking gel. Uncut vector should<br />

have three bands representing supercoiled, nicked circular, and (sometimes) linear forms.<br />

Cut vector should only have the linear form. Even small amounts of uncut vector will lead<br />

to high backgrounds on nonrecombinants.<br />

6. Run the digested material on a preparative agarose gel and cut out the vector and insert<br />

bands. Minimize exposure to ultraviolet light.<br />

7. Recover the DNA separately from each band. Many protocols are available for this: We<br />

usually use GeneClean for inserts larger than 250 bp.<br />

8. Run about one-tenth to one-fifth of the DNA extracted from the band on a checking gel<br />

to roughly estimate the amount of DNA recovered. The relative amounts of DNA in each<br />

band are crudely estimated by visually comparing the brightness of the bands to those with<br />

a known amount of DNA. The vector bands should contain more or less known quantities<br />

of DNA, if you know how much you started with, and assume about 70% was recovered<br />

from the preparative gel.<br />

9. Mix vector and insert in at least a 21 molar ratio and ligate. Use about 100 ng of vector<br />

per ligation. The ligation should resemble the following (for a larger insert, more DNA<br />

is needed for the same molar ratio): 100 ng vector (2.5 kb), 20 ng insert (250 bp), 2 µL<br />

of 10× ligation buffer, 1 µL of 10 mM ATP (if not in buffer), 0.5 µL of T4 DNA ligase,<br />

and up to 16.5 µL of H 2 O.<br />

10. Dilute the ligation 1:5, and use 1 µL to transform 20 µL of competent E. coli. Plate on<br />

appropriate antibiotic medium.<br />

11. Recombinant colonies can be screened using PCR. Depending on which restriction site<br />

was used, you may be able to screen with a T-linker primer. For example, with inserts<br />

cloned nondirectionally using the EcoRI site of the NdeT-linker, NdeTL-A can be used for<br />

screening. Make a master mix for as many reactions as you need, in which a 10-µL reaction<br />

contains: 1 µL of 10× PCR buffer, 1 µL of dNTPs (2 mM each), 1.5 µL of MgCl 2 (10 mM ),<br />

2 µL of red sucrose dye, 1 µL of NdeTL-A (10 µM ), and 0.25 µL of Taq polymerase. Cover<br />

these small reactions with oil while picking colonies to prevent evaporation.<br />

12. Touch a recombinant colony with the tip of a sterile toothpick, dip the end of the toothpick<br />

through the oil into the reaction, and swirl. Do not add enough bacteria to the reaction<br />

to make it cloudy. Just a few bacteria are sufficient. Too much bacterial matter, or


T-Linker Strategy 481<br />

small amounts of agar, will inhibit the reaction. Be sure to include one reaction of a<br />

nonrecombinant (“blue”) colony as a negative control.<br />

13. Heat to 94°C for 1 min.<br />

14. Amplify for 30 cycles using the following parameters: 94°C for 15 s, 55°C for 15 s,<br />

and 72°C for 45 s.<br />

4. Notes<br />

1. Why not use a T-vector? The method of choice for routine cloning of PCR products is<br />

probably a T-vector. One of the more important considerations is that with a T-vector you<br />

do not need to digest the insert with a restriction enzyme, so you do not need to worry about<br />

whether or not the insert contains that site. Also, reamplification with the T-linker provides<br />

another set of opportunities for the polymerase to introduce errors. However, T-linkers have<br />

several differences that can provide advantage in certain circumstances.<br />

a. The efficiency of ligation can theoretically be increased because much higher concentrations<br />

of linkers can be achieved.<br />

b. The efficiency of ligation does not need to be as high because the ligated product can<br />

be reamplified with one of the linker oligonucleotides to give a product that has added<br />

restriction sites at the ends.<br />

c. Oligonucleotides, such as those used to construct the T-linker, are quite stable, and<br />

remain usable for many years. Stability has been a problem with some of the commercially<br />

available T-vectors.<br />

d. Because the ends of a PCR product can be precisely defined and/or modified, and<br />

because T-linkers can be custom-made, it is possible to “split” a DNA sequence, such<br />

as a restriction site, between the PCR product and the T-linker so that a complete site<br />

is formed only at one end of the final product, without having to include the entire<br />

site in the primer made for amplifying a specific gene. This can save on the cost of<br />

oligonucleotides but still allow directional cloning.<br />

e. T-linkers should be more flexible in terms of using a variety of restriction sites in a<br />

variety of vectors.<br />

2. The directional cloning of the coding sequence for the α3 acetylcholine receptor subunit<br />

illustrated in Fig. 2 was accomplished by selecting the locations of the primers so that a<br />

HindIII site was created at only one end. However, by random chance, in one case of four,<br />

a PCR product made with primers not designed to complete the NdeI site will have a g at<br />

a given end, and will thus end up with an NdeI site. Similarly, one of 16 randomly chosen<br />

products will have NdeI sites at both ends. Thus, 3 of 16 randomly chosen PCR products<br />

will have a g at only one end, and could be cloned directionally using this T linker. A set<br />

of such linkers, with each depending on the presence of a different single nucleotide<br />

at the end of the PCR product, could therefore be used to directionally clone 75% of<br />

randomly chosen PCR products. Potential restriction enzymes for such complementable<br />

sites in T-linkers would be:<br />

Site ends in<br />

Restriction enzymes<br />

. . . t(g) NdeI, PvuII (blunt), Pm1I (blunt)<br />

. . . t(c) EcoRI, EcoRV (blunt), AatII, SacI<br />

. . . t(a) SnaBI (blunt)<br />

. . . t(t) Hind III, SspI (blunt)<br />

Three linkers would make up a complete set, that is, you do not need an a linker because<br />

any product that only has a at one end automatically has one of the other three bases at the


482 Horton, Raju, and Conti-Fine<br />

other end. Together with the T-linkers using split NdeI and HindIII sites described here, a<br />

T-linker that introduces a split SacI site, for example, would complete a set.<br />

3. Potentially, other DNA sequences, such as a promoter for in vitro transcription, could be<br />

split so that a functional site is completed at only one end of the product. If the majority<br />

of the DNA sequence of such sites can be added by ligating a T-linker, this could provide<br />

significant cost savings compared to synthesizing target-specific primers containing such<br />

sequences.<br />

4. Blunt-ended ligation of linkers has been used to add primer sequences to DNA fragments<br />

(9,10), but the T-linker application presented here is novel as far as we can tell. The<br />

T-linkers are more suitable for use with DNA fragments generated by PCR than bluntended<br />

linkers, because of the nontemplate derived as added by the polymerase. Because<br />

the sequences at the ends of PCR products can be easily manipulated by incorporating<br />

changes in the primers, DNA sequences, such as restriction sites, can be split between<br />

the T-linker and the PCR product. In general, this sort of site splitting is not practical<br />

except with PCR-generated fragments. One exception is fragments tailed with a known<br />

homopolymer, such as the poly A tail on eukaryotic mRNAs (11); such a linker is commercially<br />

available (Novagen, Madison, WI). Because the T-linker is added as an inverted<br />

repeat at the ends of the fragment, a single primer (TL-A) is used to reamplify after<br />

ligation. Single-primer amplification systems (12) have an advantage in that a “primerdimer”<br />

cannot be formed from one primer, because this would produce an unamplifiable<br />

hairpin.<br />

5. Because each product being cloned is subjected to an extra amplification, the overall<br />

frequency of errors should be increased, although our experience shows that the increase<br />

is not dramatic. In many cases, the risk of PCR errors is quite acceptable. For example, if<br />

one is producing an enzyme or other protein with a measurable function, the clones can be<br />

screened for activity. Deleterious mutations will thus be weeded out, and nondeleterious<br />

mutations may not matter (the α3 subunit in our example will be used as a potential<br />

ligand-binding protein, and clones will be screened on this basis). Similarly, if a sequence<br />

is to be used as a hybridization probe, a low frequency of base substitutions is frequently<br />

acceptable. In situations where no mutations are acceptable, clones must be screened<br />

by careful sequencing.<br />

6. The pH and the magnesium concentration of the PCR are different from those of the<br />

ligation. Using a small ligation volume and a large PCR volume helps to correct the<br />

conditions for the PCR. Alternatively, the pH of the bottom mix can be increased, and<br />

the added magnesium reduced, so that the final pH and (Mg 2+ ) are correct after mixing.<br />

This allows use of smaller volumes. The ability to make a quick, automated one-tube<br />

ligation/reamplification reaction makes this setup intriguing. However, it is easier to<br />

repeat parts of the experiment (such as the reamplification) if you have leftovers from<br />

a separate ligation reaction.<br />

Acknowledgments<br />

This work was supported by research grants from the Muscular Dystrophy Association<br />

of America and from the Council for Tobacco Research, by the NIH grant N52319,<br />

and the NIDA Program Project grant DA05695. R. M. H. was the recipient of the<br />

Robert G. Sampson Neuromuscular Disease Research Fellowship from the Muscular<br />

Dystrophy Association.<br />

References<br />

1. Clark, J. M. (1988) Novel nontemplated nucleotide addition reactions catalyzed by<br />

procaryotic and eucaryotic DNA polymerases. Nucleic Acids Res. 16, 9677–9686.


T-Linker Strategy 483<br />

2. Mole, S. E., Iggo, R. D., and Lane, D. P. (1989) Using the polymerase chain reaction to<br />

modify expression plasmids for epitope mapping. Nucleic Acids Res. 17, 3319.<br />

3. Mead, D. A., Pey, N. K., Herrnstadt, C., Marcil, R. A., and Smith, L. M. (1991) A universal<br />

method for the direct cloning of PCR amplified nucleic acid. <strong>Bio</strong>/Technology 9, 657–663.<br />

4. Marchuk, D., Drumm, M., Saulino, A., and Collins, F. S. (1991) Construction of T-vectors,<br />

a rapid and general system for direct cloning of unmodified PCR products. Nucleic Acids<br />

Res. 19, 1154.<br />

5. Holton, T. A. and Graham, M. W. (1991) A simple and efficient method for direct cloning<br />

of PCR products using ddT-tailed vectors. Nucleic Acids Res. 19, 1156.<br />

6. Hoppe, B. L., Conti-Tronconi, B. M., and Horton, R. H. (1992) Gel loading dyes compatible<br />

with PCR. <strong>Bio</strong>Techniques 12, 679–680.<br />

7. Horton, R. M., Hoppe, B. L., and Conti-Tronconi, B. M. (1994) AmpliGrease, “Hot Start”<br />

PCR using petroleum jelly. <strong>Bio</strong>Techniques 16, 42– 43.<br />

8. Tabor, S. and Richardson, C. C. (1985) A bacteriophage T7 RNA polymerase/promoter<br />

system for controlled exclusive expression of specific genes. Proc. Natl. Acad. Sci. USA<br />

82, 1074–1078.<br />

9. Ko, M. S. H., Takahashi, N., Nishiguchi, K., and Abe, K. (1990) Unbiased amplification<br />

of a highly complex mixture of DNA fragments by ‘lone-linker’-tagged PCR. Nucleic<br />

Acids Res. 18, 4293.<br />

10. Bhat, G. J., Lodes, M. J., Myler, P. J., and Stuart, K. D. (1991) A simple method for cloning<br />

blunt ended DNA fragments. Nucleic Acids Res. 19, 398.<br />

11. Meissner, P. S. Sisk, W. P., and Berman, M. L. (1987) Bacteriophage lambda cloning<br />

system for the construction of directional cDNA libraries. Proc. Natl. Acad. Sci. USA<br />

84, 4171– 4175.<br />

12. Rich, J. J. and Willis, D. K. (1990) A single oligonucleotide can be used to rapidly isolate<br />

DNA sequences flanking a transposon Tn5 insertion by the polymerase chain reaction.<br />

Nucleic Acids Res. 18, 6673–6676.


484 Horton, Raju, and Conti-Fine


Cloning Gene Family Members 485<br />

67<br />

Cloning Gene Family Members Using PCR<br />

with Degenerate Oligonucleotide Primers<br />

Gregory M. Preston<br />

1. Introduction<br />

1.1. What Are Gene Families?<br />

As more and more genes are cloned and sequenced, it is apparent that nearly all genes<br />

are related to other genes. Similar genes are grouped into families, such as the collagen<br />

and globin gene families. There are also gene superfamilies. Gene superfamilies are<br />

composed of genes that have areas of high homology and areas of high divergence.<br />

Examples of gene superfamilies include the oncogenes, homeotic genes, and myosin<br />

genes. In most cases, the different members of a gene family carry out related functions.<br />

A detailed protocol for the cloning by degenerate oligonucleotide polymerase chain<br />

reaction (PCR) of members of the Aquaporin family of membrane water channels<br />

(1,2) is discussed here.<br />

1.2. Advantages of PCR Cloning of Gene Family Members<br />

There are several considerations that must be taken into account when determining<br />

the advantages of using PCR to identify members of a gene family over conventional<br />

cloning methods of screening a library with a related cDNA, a degenerate primer, or<br />

an antibody. It is recommended that after a clone is obtained by PCR, one uses this<br />

template to isolate the corresponding cDNA from a library because mutations can often<br />

be introduced in PCR cloning. Alternatively, sequencing two or more PCR clones from<br />

independent reactions will also meet this objective. The following is a list of some of<br />

the advantages of cloning gene family members by PCR.<br />

1. Either one or two degenerate primers can be used in PCR cloning. When only one of the<br />

primers is degenerate, the other primer must be homologous to sequences in the phage or<br />

bacteriophage cloning vector (3,4) or to a synthetic linker sequence, as with RACE PCR.<br />

The advantage to using only one degenerate primer is that the resulting clones contain<br />

all of the genetic sequence downstream from the primer (be it 5′ or 3′ sequence). The<br />

disadvantage to this anchor PCR approach is that one of the primers is recognized by every<br />

gene in the starting material, resulting in single-strand amplification of all sequences.<br />

This is particularly notable when attempting to clone genes that are not abundant in the<br />

starting material. This disadvantage can often be ameliorated in part by using a nested<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

485


486 Preston<br />

amplification approach with two degenerate primers to preferentially amplify desired<br />

sequences.<br />

2. It is possible to perform a PCR on first-strand cDNAs made from a small amount of<br />

RNA and, in theory, from a single cell. Several single-stranded “minilibraries” can be<br />

rapidly prepared and analyzed by PCR from a number of tissues at different stages of<br />

development or cell cultures under different hormonal conditions. Therefore, PCR cloning<br />

can potentially provide <strong>info</strong>rmation about the timing of expression of an extremely rare<br />

gene family member, or messenger RNA splicing variants, that may not be present in<br />

a recombinant library.<br />

3. Finally, the time and expense required to clone a gene should be considered. Relative<br />

to conventional cloning methods, PCR cloning can be more rapid, less expensive, and<br />

in some cases, the only feasible cloning strategy. It takes at least 4 d to screen 300,000<br />

plaques from a λgt10 library. With PCR, an entire library containing 10 8 independent<br />

recombinants (~5.4 ng DNA) can be screened in one reaction. Again, to ensure authenticity<br />

of your PCR clones, you should either use the initial PCR clone to isolate a cDNA clone<br />

from a library or sequence at least two clones from independent PCRs.<br />

1.3. Degenerate Oligonucleotide Theory and Codon Usage<br />

Because the genetic code is degenerate, primers targeted to particular amino acid<br />

sequences must also be degenerate to encode the possible permutations in that sequence.<br />

Thus, a primer to a six- amino acid sequence that has 64 possible permutations can<br />

potentially recognize 64 different nucleotide sequences, one of which is to the target<br />

gene. If two such primers are used in a PCR, then there are 64 × 64 or 4096 possible<br />

permutations. The target DNA will be recognized by a small fraction (1/64) of both<br />

primers, and the amplification product from that gene will increase exponentially.<br />

However, some of the other 4095 possible permutations may recognize other gene<br />

products. This disadvantage can be ameliorated by performing nested amplifications<br />

and by using “guessmer” primers. A guessmer primer is made by considering the<br />

preferential codon usage exhibited by many species and tissues (see Subheading 3.1.).<br />

For instance, the four codons for alanine begin with GC. In the third position of<br />

this codon, G is rarely used in humans (~10.3% of the time) or rats (~8.0%), but<br />

often used in Escherichia coli (~35%) (5). This characteristic of codon usage may be<br />

advantageously used when designing degenerate oligonucleotide primers.<br />

1.4. Strategy for Cloning Aquaporin Gene Family Members<br />

In a related methods chapter (3), I described the cloning by degenerate primer PCR<br />

of Aquaporin-1 (formerly CHIP28) from a human fetal liver λgt11 cDNA library<br />

starting with the first 35 amino acids from the N-terminus of the purified protein. A<br />

full-length cDNA was subsequently isolated from an adult human bone marrow cDNA<br />

library (4) and following expression in Xenopus shown to encode a water selective<br />

channel (6). We now know that the Aquaporin family of molecular water channels<br />

includes genes expressed in diverse species, including bacteria, yeast, plants, insects,<br />

amphibians, and mammals (1,2,7). We have used degenerate oligonucleotide primers<br />

designed to highly conserved amino acids between the different members of the<br />

Aquaporin family to clone novel Aquaporin gene family cDNAs from rat brain (AQP4)<br />

and salivary gland (AQP5) libraries (8,9). In Subheading 3., I will describe the creation<br />

of a new set of degenerate primers that we have used to clone, by degenerate primer<br />

PCR, Aquaporin homologs from a number of different tissues and species. Subheading 3.


Cloning Gene Family Members 487<br />

Table 1<br />

The Degenerate Nucleotide Alphabet<br />

Letter<br />

Specification<br />

A<br />

Adenosine<br />

C<br />

Cytidine<br />

G<br />

Guanosine<br />

T<br />

Thymidine<br />

R puRine (A or G)<br />

Y pYrimidine (C or T)<br />

K Keto (G or T)<br />

M aMino (A or C)<br />

S Strong (G or C)<br />

W Weak (A or T)<br />

B Not A (G, C, or T)<br />

D Not C (A, G, or T)<br />

H Not G (A, C, or T)<br />

V Not T (A, C, or G)<br />

N aNy (A, G, C, or T)<br />

I<br />

Inosine a<br />

a Although inosine is not a true nucleotide, it is included<br />

in this degenerate nucleotide list because many researchers<br />

have employed inosine-containing oligonucleotide primers<br />

in cloning gene family members.<br />

has been broken up into three parts: Subheading 3.1. describes the designing of the<br />

degenerate primers; Subheading 3.2. describes the PCR amplification with degenerate<br />

primers; and 3. Subheading 3.3. describes the subcloning and DNA sequencing of the specific<br />

PCR-amplified products.<br />

2. Materials<br />

2.1. Design of Degenerate Oligonucleotide Primers<br />

No special materials are required here, except the amino acid sequence to which<br />

the degenerate primers will be designed and a codon usage table (5). If the degenerate<br />

primers are going to be designed according to a family of related amino acid sequences,<br />

these sequences should be aligned using a multiple sequence alignment program. A<br />

degenerate nucleotide alphabet (Table 1) provides a single-letter designation for any<br />

combination of nucleotides. Some investigators have successfully used mixed primers<br />

containing inosine where degeneracy was maximal, assuming inosine is neutral with<br />

respect to base pairing, to amplify rare cDNAs by PCR (10,11).<br />

2.2. PCR Amplification with Degenerate Primers<br />

For all buffers and reagents, distilled deionized water should be used. All buffers<br />

and reagents for PCR should be made up in distilled deionized 0.2-µ filtered water that<br />

has been autoclaved (PCR water) using sterile tubes and aerosol blocking pipet tips to<br />

prevent DNA contamination (see Note 1). All plastic supplies (microfuge tubes, pipet<br />

tips, and so on) should be sterilized by autoclaving or purchased sterile.


488 Preston<br />

1. 10× PCR buffer: 100 mM Tris-HCl, pH 8.3; at 25°C, 500 mM KCl, 15 mM MgCl 2 ; and<br />

0.1% w/v gelatin. Incubate at 50°C to melt the gelatin, filter sterilize, and store at –20°C<br />

(see Note 2).<br />

2. dNTP stock solution (1.25 mM each of dATP, dGTP, dCTP, and dTTP) made by diluting<br />

commercially available deoxynucleotides with PCR water.<br />

3. Thermostable DNA polymerase, such as Amplitaq DNA Polymerase (Perkin–Elmer Cetus,<br />

Norwalk, CT) supplied at 5 U/µL.<br />

4. Mineral oil.<br />

5. A programmable thermal cycler machine, available from a number of manufacturers,<br />

including Perkin–Elmer Cetus, MJ Research, and Stratagene.<br />

6. Degenerate oligonucleotide primers should be purified by reverse-phase high-performance<br />

liquid chromatography or elution from acrylamide gels, dried down, resuspended at<br />

20 pmol/µL in PCR-water, and stored at –20°C, preferably in aliquots.<br />

7. The DNA template can be almost any DNA sample, including a single-stranded cDNA<br />

from a reverse transcription reaction, DNA from a phage library, and genomic DNA. The<br />

DNA is heat denatured at 99°C for 10 min and stored at 4 or –20°C.<br />

8. Chloroform (see Note 3).<br />

9. Tris-saturated phenol (see Note 3), prepared using ultra-pure redistilled crystalline phenol<br />

as recommended by the supplier (Gibco-BRL [product #5509], Gaithersburg, MD). Use<br />

polypropylene or glass tubes for preparation and storage.<br />

10. PC9 (see Note 3): Mix equal volumes of buffer-saturated phenol, pH >7.2, and chloroform,<br />

extract twice with an equal volume of 100 mM Tris-HCl, pH 9.0, separate phases by<br />

centrifugation at room temperature for 5 min at 2000g, and store at 4 to –20°C for up<br />

to 1 mo.<br />

11. AmAc (7.5 M) for precipitation of DNA. Ammonium acetate is preferred over sodium<br />

acetate because nucleotides and primers generally do not precipitate with it. Dissolve in<br />

water, filter through 0.2-µm membrane, and store at room temperature.<br />

12. 100% ethanol stored at –20°C.<br />

13. 70% ethanol stored at –20°C.<br />

14. TE: 10 mM Tris, 0.2 mM EDTA, pH 8.0. Dissolve in water, filter through 0.2-µm<br />

membrane, and store at room temperature.<br />

15. 50× TAE: 242 g of Tris-HCl base, 57.1 mL of acetic acid, 18.6 g of Na 2 (H 2 O) 2 EDTA.<br />

Dissolve in water, adjust volume to 1 L, and filter through 0.2-µm membrane. Store at<br />

room temperature.<br />

16. HaeIII-digested φX174 DNA markers. Other DNA molecular weight markers can be used<br />

depending on availability and the size of the expected PCR-amplified products.<br />

17. 6× gel loading buffer (GLOB): 0.25% bromophenol blue, 0.25% xylene cyanol FF, 1 mM<br />

EDTA, and 30% glycerol in water. Store up to 4 mo at 4°C.<br />

18. Agarose gel electrophoresis apparatus and electrophoresis grade agarose. For the optimal<br />

resolution of DNA products >500 bp in length, NuSieve GTG agarose (FMC <strong>Bio</strong>Products)<br />

is recommended.<br />

19. Ethidium bromide (EtBr; see Note 3). Ethidium bromide (10 mg/mL stock) prepared in<br />

water and stored at 4°C in a brown or foil-wrapped bottle. Use at 0.5 to 2.0 µg/mL in water<br />

for staining nucleic acids in agarose or acrylamide gels.<br />

20. For the elution of specific PCR-amplified DNA products from agarose gels, several methods<br />

are available, including electroelution and electrophoresis onto DEAE-cellulose membranes<br />

(12,13). Several commercially available kits will also accomplish this task. I have had<br />

some success with GeneClean II (<strong>Bio</strong> 101, La Jolla, CA) for PCR products >500 bp<br />

in length, and with QIAEX (Qiagen, Chatsworth, CA) for products from 50 to 5000 bp. If<br />

you do not know the approximate size of the PCR-amplified products and wish to clone


Cloning Gene Family Members 489<br />

all of them, the QIAquick-spin PCR purification kit is recommended (Qiagen) because<br />

this will remove all nucleotides and primers before attempting to clone. This kit is<br />

also recommended for purification of PCR products for secondary PCR-amplification<br />

reactions.<br />

2.3. Cloning and DNA Sequencing of PCR-Amplified Products<br />

1. From Subheading 2.2., items 8–14 and 20.<br />

2. pBluescript II phagemid vector (Stratagene). A number of comparable bacterial expression<br />

vectors are available from several companies.<br />

3. Restriction enzymes: EcoRV (for blunt-end ligation).<br />

4. Calf intestinal alkaline phosphatase (CIP; New England <strong>Bio</strong>labs, Beverly, MA).<br />

5. Klenow fragment of Escherichia coli DNA polymerase I (sequencing grade preferred) and<br />

10 mM dNTP solution (dilute PCR or sequencing grade dNTPs).<br />

6. T4 DNA ligase (1 or 5 U/µL) and 5× T4 DNA ligase buffer (Gibco-BRL).<br />

7. Competent DH5α bacteria. Can be prepared (12,13) or purchased. Other bacterial strains<br />

can be substituted.<br />

8. Ampicillin: 50 mg/mL stock in water, 0.2-µ filtered, and stored in aliquots at –20°C<br />

(see Note 4).<br />

9. LB media: 10 g of bacto-tryptone, 5 g of bacto-yeast extract, and 10 g of NaCl dissolved in<br />

1 L of water. Adjust pH to 7.0. Sterilize by autoclaving for 20 min on liquid cycle.<br />

10. LB-Amp plates: Add 15 g of bacto-agar to 1000 mL of LB media before autoclaving for<br />

20 min on the liquid cycle. Gently swirl the media on removing it from the autoclave to<br />

distribute the melted agar. Be careful: The fluid may be superheated and may boil over<br />

when swirled. Place the media in a 50°C water bath to cool. Add 1 mL of ampicillin,<br />

swirl to distribute, and pour 25 to 35 mL/90-mm plate. Carefully flame the surface of the<br />

media with a Bunsen burner to remove air bubbles before the agar hardens. Store inverted<br />

overnight at room temperature, then wrapped at 4°C for up to 6 mo.<br />

11. IPTG: Dissolve 1 g of isopropylthiogalactoside in 4 mL of water, filter through 0.2-µm<br />

membrane, and store in aliquots at –20°C.<br />

12. X-Gal: Dissolve 100 mg of 5-bromo-4-chloro-3-indolyl-β-D-galactopyranoside in 5 mL of<br />

dimethylformamide and stored at –20°C in a foil wrapped tube (light sensitive).<br />

13. Plasmid DNA isolation equipment and supplies (12,13) or plasmid DNA isolation kits,<br />

available from many manufacturers.<br />

14. Double-stranded DNA sequencing equipment and supplies (12,13) or access to a DNA<br />

sequencing core facility.<br />

3. Methods<br />

3.1. Design of Degenerate Oligonucleotide Primers<br />

1. The first step in designing a degenerate primer is to select a conserved amino acid sequence<br />

and then determine the potential nucleotide sequence (or the complement of this sequence<br />

for a downstream primer), considering all possible permutations. If the amino acid<br />

sequence is relatively long, you can potentially design two or more degenerate primers.<br />

If only one is made, make it to sequences with a high (50–65%) GC content because<br />

these primers can be annealed under more stringent conditions (e.g., higher temperatures).<br />

Figure 1 shows an alignment of the amino acid sequences for several members of the<br />

Aquaporin gene family in the two most highly conserved regions. Also shown is the<br />

consensus amino acid sequence, the degenerate nucleotide sequence, and the sequence<br />

of the primers we used to isolate Aquaporin gene family members. Interestingly, not<br />

only are these two regions highly conserved among the different members of this gene


490 Preston<br />

Fig. 1. Design of degenerate primers to amplify Aquaporin gene family members. Top, the<br />

amino acid sequences of 10 MIP family proteins, including the S. cerevisiae FPS1 (23), E.<br />

coli GlpF (24), α- and γ-tonoplast intrinsic proteins (TIP) of Arabidopsis thaliana (25), the<br />

vasopressin-responsive water channel of rat renal collecting tubules (AQP2) (26), the major<br />

intrinsic protein (MIP) of bovine lens fiber membranes (27), human Aquaporin-1 (4), turgor<br />

responsive gene (TUR) 7a from Pisum stivum (28), the Drosophila neurogenic big brain protein<br />

(29), and the Rhyzodium root Nodulin-26 peribacteroid membrane protein (30) were aligned<br />

by the PILEUP program of progressive alignments (31) using a gap weight of 3.0 and a gap<br />

length of 0.1 running on a VAX computer system. The two most highly conserved regions are<br />

shown, separated by the number of intervening amino acids. The most highly conserved amino<br />

acids are enclosed. Middle, below the aligned sequences, the consensus amino acid sequences<br />

are shown. Bottom, from part of the consensus amino acid sequences, the degenerate nucleotide<br />

sequences were determined (using the degenerate nucleotide alphabet from Table 1) followed<br />

by the sequences for the degenerate oligonucleotide primers.<br />

family, but they are also highly related to each other, with the conserved motif being<br />

(T/S)GxxxNPAxx(F/L)G, that has been speculated to have resulted from an ancient<br />

internal duplication in a primordial bacterial organism, because this repeat has persisted<br />

in Aquaporin homologs from bacteria through plants and mammals (1,6,14). These two<br />

regions are functionally related, both contributing to the formation of the water pore in<br />

Aquaporin-1 (15).<br />

2. The next step is to determine the number of permutations in the nucleotide sequence. There<br />

are 192 permutations ([2 × 4] × 3 × 4 × 2) in the sequence 5′-YTN-ATH-GGN-GAR-3′,<br />

which encodes the hypothetical amino acid sequence Leu-Ile-Gly-Glu. We can reduce the<br />

degeneracy by making educated guesses in the nucleotide sequence, that is, by making a<br />

guessmer. The 3′-end of a primer should contain all possible permutations in the amino<br />

acid sequence because Taq DNA polymerase will not extend a prime with a mismatch at the<br />

extending (3′) end. If the above primer was to a human gene, a potential guessmer would<br />

be 5′-CTB-ATY-GGN-GAR-3′, which only contains 64 permutations. This guessmer is<br />

proposed by taking into account the preferential codon usage for leucine and isoleucine<br />

in humans (5).<br />

3. The degeneracy of a primer can be reduced further by incorporating inosine residues in the<br />

place of N. The advantages of using inosine-containing primers is that they have a reduced<br />

number of permutations and that the inosine reportedly base pairs equally well with all


Cloning Gene Family Members 491<br />

four nucleotides, creating a single bond in all cases (10). The disadvantage is that inosines<br />

reduce the annealing temperature of the primer. I have not used inosine-containing primers<br />

in my studies.<br />

4. It is often convenient to incorporate restriction endonuclease sites at the 5′-ends of a primer<br />

to facilitate cloning into plasmid vectors (4,8,9). Different restriction sites can be added<br />

to the 5′-ends of different primers so the products can be cloned directionally. However,<br />

not all restriction enzymes can recognize cognate sites at the ends of a double-stranded<br />

DNA molecule equally well. This difficulty can often be reduced by adding a two to four<br />

nucleotide 5′-overhang before the beginning of the restriction enzyme site (see Note 5).<br />

Some of the best restriction enzymes sites to use are EcoRI, BamHI, and XbaI. Catalogs<br />

from New England <strong>Bio</strong>labs have a list of the ability of different restriction enzymes to<br />

recognize short base-pair sequences. A potential pitfall of this approach would be the<br />

occurrence of the same restriction site within the amplified product as used on the end of<br />

one of the primers. Therefore, only part of the amplified product would be cloned.<br />

5. The final consideration you should make is the identity of the 3′ most nucleotide. The<br />

nucleotide on the 3′-end of a primer should preferably be G or C and not N, I, or T.<br />

The reason for this is that thymidine (and supposedly inosine) can nonspecifically prime<br />

on any sequence. Guanosines and cytidine are preferred since they form three H-bonds at<br />

the end of the primer, a degree stronger than an AT base pair.<br />

3.2. PCR Amplification and DNA Purification<br />

The template for these reactions can be the DNA in a phage library or the first-strand<br />

cDNA from a reverse transcription reaction on RNA. A phage library with a titer of<br />

5 × 10 9 pfu/mL would contain, in a 5-µL aliquot, 2.5 × 10 7 pfu (~1.5 ng of DNA).<br />

Before PCR amplification, the DNA is heat denatured at 99°C for 10 min.<br />

3.2.1. PCR (see Notes 1 and 6)<br />

In all cases, the DNA template should also be PCR amplified with the individual<br />

degenerate primers to determine whether any of the bands amplified are derived from<br />

one of the degenerate primer pools. A DNA-free control is required to assess if there is<br />

contaminating DNA in any of the other reagents.<br />

1. Pipet into 0.5-mL microcentrifuge tubes in the following order: 58.5 µL of PCR-water that<br />

has been autoclaved; 10 µL of 10× PCR buffer (see Note 2); 16 µL of 1.25 mM dNTP stock<br />

solution; 5.0 µL of primer up-1; 5.0 µL of primer down-1; and 5.0 µL of heat-denatured<br />

library or cDNA (1–100 ng). If several reactions are being set up concurrently, a master<br />

reaction mix can be made up consisting of all the reagents used in all of the reactions, such<br />

as the PCR water, reaction buffer, and dNTPs.<br />

2. Briefly vortex each sample and spin for 10 s in a microfuge. Overlay each sample with<br />

2 to 3 drops of mineral oil.<br />

3. Amplify by hot-start PCR using the following cycle parameters. Pause the thermocycler<br />

in step 4-cycle 1 and add 0.5 µL of Amplitaq DNA polymerase to each tube. Then, run for<br />

95°C, 5 min (initial denaturation); 94°C, 60 s (denaturation); 50°C, 90 s (annealing; see<br />

Note 7); 72°C, 60 s (extension); cycle 29 times to step 2; 72°C, 4 min; and 10°C hold.<br />

3.2.2. DNA Isolation and Gel Electrophoresis Analysis<br />

1. Remove the reaction tubes from the thermal cycler and add 200 µL of chloroform. Spin for<br />

10 s in a microfuge to separate the oil-chloroform layer from the aqueous layer. Carefully<br />

transfer the aqueous layer to a clean microfuge tube.


492 Preston<br />

2. Remove the AmpliTaq DNA polymerase by extracting the aqueous phase twice with 100 µL<br />

of PC9 (see Note 3). Spin for 2 min in a microfuge to separate the lower organic layer<br />

from the upper aqueous layer and transfer the aqueous layer to clean microfuge tube. This<br />

step is essential before digesting the DNA with restriction enzymes for directional cloning<br />

(see Subheading 3.3.) because the polymerase can precipitate, and in the presence of<br />

nucleotides, fill in recessed 3′ termini on DNA.<br />

3. AmAc-EtOH precipitation: To a 100 µL of DNA sample add 50 µL of 7.5 M AmAc (50%<br />

vol). Vortex briefly to mix. Precipitate the DNA with 350 µL of 100% ethanol (2–2.5 vol).<br />

Vortex the samples for 15 s and ice for 15 min. Spin down the DNA at 12,000g for 15 min<br />

at 4°C in a microfuge. Decant the aqueous waste. Add 250 µL of 70% ethanol. Vortex<br />

briefly and spin another 5 min at 4°C. Decant the ethanol and allow the pellets to dry<br />

inverted at room temperature, or dry in a Speed-Vac for 2 to 10 min.<br />

4. Resuspend in 20 µL of PCR water.<br />

5. The next step is to resolve an aliquot (2–10 µL) of the PCR fragments by gel electrophoresis.<br />

Small DNA products (>300 bp) can be resolved at high resolution on 5 to 10% polyacrylamide<br />

gels (12,13). Moderate-sized PCR products (150–1000 bp) should be resolved<br />

on 2 to 4% NuSieve agarose gels (in 1× TAE buffer). Larger PCR products (>500 bp)<br />

can be resolved on 0.8 to 2% agarose gels (1× TAE buffer).<br />

6. After the bromophenol blue dye has reached the end of the gel, soak the gel for 5 to 30 min<br />

in about 10 vol of water containing 1 µg/mL EtBr (see Note 3). Then view and photograph<br />

the gel under ultraviolet light. As shown in Fig. 1, there is little variability in the distance<br />

between the NPA motifs with the known members of the Aquaporin gene family. PCR<br />

amplification of the known Aquaporins cDNAs using the internal degenerate primers<br />

would generate products from 345 to 415 bp. A typical result is shown in Fig. 2.<br />

3.2.3. Secondary PCR Amplifications and DNA Purification<br />

Based on the results from gel electrophoresis of the PCR-amplified DNA products, a<br />

decision must be made on what to do next. The options are the following.<br />

1. Amplify by PCR from the initial DNA sample under different conditions.<br />

2. Amplify by PCR from a different DNA sample under the same conditions. (Different<br />

MgCl 2 concentration, annealing temperature, or primers, see Notes 2, 6, and 7).<br />

3. Gel purify a band(s) of DNA from the gel for cloning or to reamplify by PCR.<br />

4. Purify all PCR-amplified DNA fragments for cloning or to reamplify by PCR.<br />

5. Reamplify by PCR with the same or an internal pair of degenerate primers.<br />

Options 1 and 2 are self explanatory. If you want to gel purify a particular band or<br />

group of bands from an agarose gel, a number of procedures and kits are available (see<br />

Subheading 2.2.). If you plan on immediately cloning a PCR band(s), you may want<br />

to run the rest of the initial PCR on another gel to increase the recovery of DNA. It<br />

is also possible to recover specific DNA fragments from an acrylamide gel (3,12,13).<br />

To purify all PCR-amplified DNA fragments from the remaining sample, a number of<br />

methods are available, including the QIAquick-spin PCR purification kit, which can<br />

be used instead of steps 1–3 in Subheading 3.2.2. (Qiagen). Finally, aliquots of<br />

DNA purified from a gel or from the initial PCR (1–10%) can be reamplified by PCR<br />

with either the same or an internal pair of degenerate oligonucleotide primers (see<br />

Note 1).<br />

When attempting to identify a gene family homolog from a tissue that is known to<br />

express a homolog(s), a number of tricks can be tried to enrich the final PCR sample


Cloning Gene Family Members 493<br />

Fig. 2. Gel electrophoresis analysis of PCR-amplified DNA. DNA isolated from a human<br />

kidney cDNA library in bacteriophage λgt10 was amplified with degenerate primers up-1 (lanes<br />

1, 5, and 6), up-2 (lanes 2, 7, and 8), down-1 (lanes 3, 5, and 7), and down-2 (lanes 4, 6, and 8).<br />

Reactions containing 5 × 10 6 pfu of heat-denatured phage DNA, 100 pmol of degenerate primers,<br />

and 1.5 mM MgCl 2 in a 100-µL volume were subject to 40 cycles of PCR amplification under<br />

the following parameters: 94°C for 60 s, 48°C for 90 s, and 72°C for 60 s. After chloroform<br />

extraction and ethanol precipitation, the DNA was resuspended in 20 µL of water, and 5 µL was<br />

electrophoresed into a 4% NuSieve agarose gel in 1× TAE. The gel was stained with ethidium<br />

bromide and photographed. The relative mobility of HaeIII digested φX174 DNA markers is<br />

shown on the right. The bracket shows the size range of known members of this gene family<br />

from the primers used.<br />

for new homologs. Because the degenerate oligonucleotide primers are designed from<br />

the sequence of the known gene family members, these primers will likely be biased<br />

for those homologs. Aquaporin-1 is abundant in the capillaries around the salivary<br />

glands and throughout the body, but absent in the salivary gland (16). To identify a<br />

salivary homology of the Aquaporin gene family, we used a rat salivary gland cDNA<br />

library that also contained Aquaporin-1 cDNAs, presumably from the surrounding<br />

capillaries. We first amplified the cDNA library with an external set of degenerate<br />

primers, digested the PCR-amplified DNAs with the restriction enzyme PstI (which<br />

cuts between the NPA motifs of rat AQP1), and reamplified with an internal pair of<br />

primers. We again digested with PstI to digest the rat AQP1 DNAs, then cloned and<br />

sequenced the DNA fragments between 350 and 450 bp (9). This strategy would not<br />

work if the resulting cDNA (AQP5) also contained a PstI site. By trying different<br />

restriction enzymes that cut DNA infrequently (6–8 bp-recognition sites), a number<br />

of new homologs will preferentially be identified. Alternatively, after cloning the<br />

DNA products into bacterial expression vectors, bacterial colony lift hybridization<br />

can be used to identify colonies containing inserts for known gene family members<br />

(3,12,13).


494 Preston<br />

3.3. Cloning and DNA Sequencing of PCR-Amplified Products<br />

3.3.1. Preparation of Vector for Ligation<br />

1. For blunt-end ligations, digest 1 µg of pBluescript II KS phagemid DNA (Stratagene) with<br />

10 U EcoRV in a 50-µL vol. Incubate at 37°C for 2 h. For cohesive-end ligations, similarly<br />

digest the vector with the appropriate restriction enzyme(s).<br />

2. For both blunt-end ligations and cohesive-end ligations where the vector has been digested<br />

with only one restriction enzyme, it is necessary to remove the 5′-phosphate from the<br />

vector to inhibit the vector from self ligating. This is accomplished by treating the vector<br />

with CIP according to the manufacturer’s recommendations. Note that 1 µg of a 3-kbp<br />

linear DNA molecule contains 1 pmol of 5′ overhangs (BamHI), blunt-ends (EcoRV), or<br />

3′-overhangs (PstI), depending on the enzyme that digested it. Afterward, add EDTA to<br />

5 mM and heat-kill the enzyme at 65°C for 1 h. Adjust the volume to 50 to 100 µL with TE<br />

and extract once with Tris-saturated phenol, twice with PC9, and twice with chloroform.<br />

Back extract each organic layer with 50 µL of TE and pool with the final sample.<br />

AmAc-EtOH precipitate (see Subheading 3.2.2.) and resuspend in 10 µL of water.<br />

3. If the insert is going to be directionally cloned into the vector, just extract once with<br />

50 µL of PC9, AmAc-EtOH precipitate (see Subheading 3.2.2.) and resuspend in 10 µL<br />

of water.<br />

3.3.2. Preparation of Inserts for Ligation<br />

AmpliTaq and other thermostable DNA polymerases often fail to completely fill in<br />

the ends of the double-stranded DNA products, thus leaving recessed 3′ termini that<br />

can be filled in with the Klenow fragment of E. coli DNA polymerase I. This should be<br />

done whether or not the DNA is going to be digested with restriction enzymes added to<br />

the ends of the primers for directional cloning (see Subheading 3.1.).<br />

1. AmAc-EtOH precipitate the DNA (see Subheading 3.2.2.) and resuspend in 15 µL of<br />

water.<br />

2. Add 2 µL of 10× restriction enzyme reaction buffer. Klenow DNA polymerase works<br />

well in most restriction enzyme digestion buffers (10× REact 2 or 3 from Gibco-BRL).<br />

If the DNA is going to be subsequently digested with a restriction enzyme(s), use the<br />

buffer for that enzyme.<br />

3. Add 2 µL of 10 mM dNTP solution. Then, add Klenow DNA polymerase (1 U/µg DNA)<br />

and incubate at room temperature for 15 min.<br />

4. Heat-inactivate the enzyme at 75°C for 10 min. If the DNA is going to be directly used in<br />

ligation reactions, it is not necessary to purify the DNA from the unincorporated dNTPs<br />

because they will not inhibit T4 DNA ligase. To concentrate the DNA sample, proceed<br />

with step 6.<br />

5. PCR products containing restriction sites on their ends should now be digested with the<br />

restriction enzymes. Incubate in the appropriate buffer, using 20 U of enzyme/µg of DNA<br />

and incubating for 2 to 4 h at the proper temperature.<br />

6. Extract the DNA once or twice with PC9 and precipitate with AmAc-EtOH as described<br />

above (see Subheading 3.2.2.). Resuspend the final pellet in 5 to 10 µL of water.<br />

3.3.3. DNA Ligation and Bacterial Transformation<br />

1. At this point, it is often advantageous to run a small aliquot of the different DNA fragments<br />

on a gel to assess their approximate concentrations and purity. Ideally, you want at least a<br />

21 molar ratio of insert to vector in the ligation reactions. If necessary return to the above<br />

procedures to isolate more DNA for the ligation reaction.


Cloning Gene Family Members 495<br />

2. Set up the ligation reactions with the vector and insert similar to the following:<br />

a. Reaction 1: 1 µL of vector (10 ng; vector control);<br />

b. Reaction 2: 1 µL of vector + 1 µL of insert (~10 ng insert);<br />

c. Reaction 3: 1 µL of vector + 4 µL of insert.<br />

Then add 2 µL of 5× T4 DNA ligase buffer (Gibco-BRL) and water to 9.5 µL. If the<br />

buffer is more than 4-mo-old, the ATP may be depleted. Therefore, add fresh ATP to a<br />

final concentration of 1 mM.<br />

3. For cohesive-end ligations add 0.5 µL of T4 DNA ligase (1 U/µL), gently mix, spin 5 s<br />

in a microfuge, and incubate at 15°C for 10 to 20 h. For blunt-end ligations add 1 µL<br />

of T4 DNA ligase (5 U/µL), gently mix, spin 5 s in a microfuge, and incubate at 25°C<br />

(or room temperature) for 1 to 12 h. Stop the reaction by heating at 75°C for 10 min and<br />

store the samples at –20°C.<br />

4. Set up a bacterial transformation with competent DH5α bacteria or a comparable strain<br />

of bacteria. Be sure to include a positive control (10 ng of undigested vector DNA) and a<br />

negative control (water). To 1.5-mL microfuge tubes, add half of the ligation mix (5 µL)<br />

or 5 µL of control DNA or water and 50 µL of competent bacteria (thawed slowly on ice);<br />

incubate on ice for 30 min. Heat-shock at 42°C for 2 min. Return to ice for 1 min. Add<br />

200 µL of LB media containing 10% glycerol. Mix gently and allow bacteria to recover<br />

and express the ampicillin resistance gene by incubating at 37°C for 1 h.<br />

5. Prewarm LB-Amp plates at 37°C for 45 min. About 30 min before plating the bacteria on<br />

the plates, add 40 µL of X-Gal and 4 µL of IPTG and quickly spread over the entire surface<br />

of the plate using a sterile glass spreader. Spread 20 to 200 µL of the transformation<br />

reactions on these plates. Allow the inoculum to absorb into the agar and incubate the<br />

plates inverted at 37°C for 12 to 24 h (see Note 4). Afterward, placing the plates at 4°C<br />

for 2 to 4 h will help enhance the blue color development.<br />

3.3.4. Plasmid DNA Minipreps and DNA Sequencing<br />

1. Colonies that contain active β-galactosidase will appear blue, whereas those containing<br />

a disrupted LacZ gene will be white. Set up minicultures by inoculating individual white<br />

colonies into 2 mL of LB media containing ampicillin. After growing at 37°C overnight,<br />

isolate the plasmid DNA. Resuspend the DNA in 20 to 50 µL of water or TE.<br />

2. Digest 5 to 20 µL of the DNA with the appropriate restriction enzymes and analyze by<br />

agarose gel electrophoresis (see Subheading 3.2.3.).<br />

3. Perform double-stranded DNA sequencing on recombinants containing inserts in the<br />

expected size range.<br />

4. Notes<br />

1. All PCRs should be set up in sterile laminar flow hoods using pipet tips containing filters<br />

(aerosol-resistant tips) to prevent the contamination of samples, primers, nucleotides,<br />

and reaction buffers by DNA. If the PCR is going to be reamplified by PCR, all possible<br />

intervening steps should also be performed in a sterile hood with the same precautions to<br />

prevent DNA contamination. These precautions should also be extended to all extractions<br />

and reactions on the nucleic acid (RNA or DNA) through the last PCR. Likewise, all<br />

primers, nucleotides, and reaction buffers for PCR should be made up and aliquoted using<br />

similar precautions. All buffers for PCR should be made with great care using sterile<br />

disposable plastic or baked glass, and restricted for use with aerosol-resistant pipet tips.<br />

2. Standard PCR buffers contain 15 mM MgCl 2 (1.5 mM final concentration). In many<br />

cases, changes in the MgCl 2 concentration will have significant consequences on the<br />

amplification of specific bands. In PCR-amplifying the four exons of the AQP1 gene,


496 Preston<br />

MgCl 2 concentrations between 0.7 and 1.0 mM gave the best results (17,18); however,<br />

MgCl 2 concentrations between 0.5 and 5.0 mM have been reported.<br />

3. Organic solvents and ethidium bromide are hazardous materials. Always handle with<br />

tremendous caution, wearing gloves and eye protection. Contact your hazardous waste<br />

department for proper disposal procedures in your area.<br />

4. Ampicillin-resistant bacteria secrete β-lactamase into the media, which rapidly inactivates<br />

the antibiotic in regions surrounding the growing bacterial colony. Thus, when bacteria are<br />

growing at a high density or for long periods (>16 h), ampicillin-sensitive satellite colonies<br />

will appear around the primary colonies (which are white in blue-white selections).<br />

This problem can be ameliorated (but not eliminated) by substitution of carbenicillin for<br />

ampicillin on agar plates.<br />

5. When designing primers with restriction enzyme sites and 5′-overhangs, note that the<br />

5′-overhang should not contain sequences complementary to the sequence just 3′ of the<br />

restriction site because this would facilitate the production of primer-dimers. Consider<br />

the primer 5′-ggg.agatct.CCCAGCTAGCTAGCT-3′, which has a XbaI site proceeded by<br />

a 5′-ggg and followed by a CCC-3′. These 12 nucleotides on the 5′-end are palindromic<br />

and can therefore easily dimerize with another like primer. A better 5′-overhang would<br />

be 5′-cac.<br />

6. When cloning a gene from a recombinant library by PCR, remember that not all genes are<br />

created equally. Genes with high GC contents have proven more difficult to clone than<br />

most. Several researchers have made contributions in a search for factors to enhance the<br />

specificity of PCRs. Nonionic detergents, such a Nonident P-40, can be incorporated in<br />

rapid sample preparations for PCR analysis without significantly affecting Taq polymerase<br />

activity (19). In some cases, such detergents are absolutely required to reproducibly detect<br />

a specific product (20) presumably because of inter- and intrastrand secondary structure.<br />

Tetramethylammonium chloride has been shown to enhance the specificity of PCRs by<br />

reducing nonspecific priming events (21). Commercially available PCR enhancers are<br />

also available.<br />

7. A critical parameter when attempting to clone by PCR is the selection of a primer annealing<br />

temperature. This is especially true when using degenerate primers. The primer melting<br />

temperature (T m ) is calculated by adding 2° for AT base pairs, 3° for GC base pairs; 2°<br />

for NN base pairs, and 1° for IN base pairs. Most PCR chapters suggest you calculate the<br />

T m and set the primer annealing temperature to 5 to 10°C below the lowest T m . Distantly<br />

related gene superfamily members have been cloned using this rationale (22). However, I<br />

have found that higher annealing temperatures are helpful in reducing nonspecific priming,<br />

which can significantly affect reactions containing degenerate primers.<br />

Acknowledgments<br />

I thank my colleagues, especially Peter Agre and William B. Guggino, for their<br />

support and helpful discussions. This work was supported in part by NIH grants<br />

HL33991 and HL48268 to Peter Agre.<br />

References<br />

1. Reizer, J., Reizer, A., and Saier, M. H., Jr. (1993) The MIP family of integral membrane<br />

channel proteins: sequence comparisons, evolutionary relationships, reconstructed pathways<br />

of evolution, and proposed functional differentiation of the two repeated halves of the<br />

proteins. Crit. Rev. <strong>Bio</strong>chem. Mol. <strong>Bio</strong>l. 28, 235–257.<br />

2. Knepper, M. A. (1994) The aquaporin family of molecular water channels. Proc. Natl.<br />

Acad. Sci. USA 91, 6255–6258.


Cloning Gene Family Members 497<br />

3. Preston, G. M. (1993) Use of degenerate oligonucleotide primers and the polymerase<br />

chain reaction to clone gene family members, in Methods in Molecular <strong>Bio</strong>logy, vol. 15:<br />

PCR Protocols: Current Methods and Applications. (White, B. A., ed.) Humana, Totowa,<br />

NJ, pp. 317–337.<br />

4. Preston, G. M. and Agre, P. (1991) Isolation of the cDNA for erythrocyte integral membrane<br />

protein of 28 kilodaltons: Member of an ancient channel family. Proc. Natl. Acad. Sci.<br />

USA 88, 11,110–11,114.<br />

5. Wada, K.-N., Aota, S.-I., Tsuchiya, R., Ishibashi, F., Gojobori, T., and Ikemura, T. (1990)<br />

Codon usage tabulated from the GenBank genetic sequence data. Nucleic Acids Res. 18,<br />

2367–2411.<br />

6. Preston, G. M., Carroll, T. P., Guggino, W. B., and Agre, P. (1992) Appearance of water<br />

channels in Xenopus oocytes expressing red cell CHIP28 protein. Science 256, 385–387.<br />

7. Chrispeels, M. J. and Agre, P. (1994) Aquaporins: Water channel proteins of plant and<br />

animal cells. TIBS 19, 421– 425.<br />

8. Jung, J. S., Bhat, B. V., Preston, G. M., Guggino, W. B., Baraban, J. M., and Agre P. (1994)<br />

Molecular characterization of an aquaporin cDNA from brain: candidate osmoreceptor and<br />

regulator of water balance. Proc. Natl. Acad. Sci. USA 91, 13,052–13,056.<br />

9. Raina, S., Preston, G. M., Guggino, W. B., and Agre, P. (1995) Molecular cloning and<br />

characterization of an aquaporin cDNA from salivary, lacrimal and respiratory tissues.<br />

J. <strong>Bio</strong>l. Chem. 270, 1908–1912.<br />

10. Knoth, K., Roberds, S., Poteet, C., and Tamkun, M. (1988) Highly degenerate, inosinecontaining<br />

primers specifically amplify rare cDNA using the polymerase chain reaction.<br />

Nucleic Acids Res. 16, 10,932.<br />

11. Chérif-Zahar, B., Bloy, C., Kim, C. L. V., Blanchard, D., Bailly, P., Hermand, P., et al.<br />

(1990) Molecular cloning and protein structure of a human blood group Rh polypeptide.<br />

Proc. Natl. Acad. Sci. USA 87, 6243–6247.<br />

12. Sambrook, J., Fritsch, E. F., and Maniatis, T., eds. (1989) Molecular Cloning: A Laboratory<br />

Manual. Cold Spring Harbor Laboratory, Cold Spring Harbor, NY.<br />

13. Ausubel, F. M., Brent, R., Kingston, R. E., Moore, D. D., Seidman, J. G., Smith, J. A., et al.,<br />

eds. (1994) Current Protocols in Molecular <strong>Bio</strong>logy. Greene Publishing/Wiley-Interscience,<br />

New York.<br />

14. Wistow, G. J., Pisano, M. M., and Chepelinsky, A. B. (1991) Tandem sequence repeats in<br />

transmembrane channel proteins. TIBS 16, 170–171.<br />

15. Jung, J. S., Preston, G. M., Smith, B. L., Guggino, W. B., and Agre, P. (1994) Molecular<br />

structure of the water channel through aquaporin CHIP: The hourglass model. J. <strong>Bio</strong>l.<br />

Chem. 269, 14,648–14,654.<br />

16. Nielsen, S., Smith, B. L., Christensen, E. I., and Agre, P. (1993) Distribution of the<br />

aquaporin CHIP in secretory and resorptive epithelia and capillary endothelia. Proc. Natl.<br />

Acad. Sci. USA 90, 7275–7279.<br />

17. Smith, B. L., Preston, G. M., Spring, F. A., Anstee, D. J., and Agre, P. (1994) Human<br />

red cell Aquaporin CHIP, I. molecular characterization of ABH and Colton blood group<br />

antigens. J. Clin. Invest. 94, 1043–1049.<br />

18. Preston, G. M., Smith, B. L., Zeidel, M. L., Moulds, J. J., and Agre, P. (1994) Mutations<br />

in aquaporin-1 in phenotypically normal humans without functional CHIP water channels.<br />

Science 265, 1585–1587.<br />

19. Weyant, R. S., Edmonds, P., and Swaminathan, B. (1990) Effects of ionic and nonionic<br />

detergents on the Taq polymerase. <strong>Bio</strong>technology 9, 308–309.<br />

20. Bookstein, R., Lai, C.-C., To, H., and Lee, W.-H. (1990) PCR-based detection of a<br />

polymorphic BamHI site in intron 1 of the human retinoblastoma (RB) gene. Nucleic<br />

Acids Res. 18, 1666.


498 Preston<br />

21. Hung, T., Mak, K., and Fong, K. (1990) A specificity enhancer for polymerase chain<br />

reaction. Nucleic Acids Res. 18, 4953.<br />

22. Zhao, Z.-Y. and Joho, R. H. (1990) Isolation of distantly related members in a multigene<br />

family using the polymerase chain reaction technique. <strong>Bio</strong>chem. <strong>Bio</strong>phys. Res. Commun.<br />

167, 174–182.<br />

23. Aelst, L. V., Hohmann, S., Zimmermann, F. K., Jans, A. W. H., and Thevelein, J. M. (1991)<br />

A yeast homologue of the bovine lens fiber MIP gene family complements the growth<br />

defect of a Saccharomyces cerevisiae mutant on fermentable sugars but not its defect in<br />

glucose-induced RAS-mediated cAMP signalling. EMBO J. 10, 2095–2104.<br />

25. Muramatsu, S. and Mizuno, T. (1989) Nucleotide sequence of the region encompassing<br />

the glpKF operon and its upstream region containing a bent DNA sequence of Escherichia<br />

coli. Nucleic Acids Res. 17, 4378.<br />

25. Höfte, H., Hubbard, L., Reizer, J., Ludevid, D., Herman, E. M., and Chrispeels, M. J.<br />

(1992) Vegetative and seed-specific forms of Tonoplast Intrinsic Protein in the vacuolar<br />

membrane of Arabidopsis thaliana. Plant Physiol. 99, 561–570.<br />

26. Fushimi, K., Uchida, S., Hara, Y., Hirata, Y., Marumo, F., and Sasaki, S. (1993) Cloning<br />

and expression of apical membrane water channel of rat kidney collecting tubule. Nature<br />

361, 549–552.<br />

27. Gorin, M. B., Yancey, S. B., Cline, J., Revel, J.-R., and Horwitz, J. (1984) The major<br />

intrinsic protein (MIP) of the bovine lens fiber membrane: characterization and structure<br />

based on cDNA cloning. Cell 39, 49–59.<br />

28. Guerrero, F. D., Jones, J. T., and Mullet, J. E. (1990) Turgor-responsive gene transcription<br />

and RNA levels increase rapidly when pea shoots are wilted: sequence and expression of<br />

three induced genes. Plant Mol. <strong>Bio</strong>l. 15, 11–26.<br />

29. Rao, Y., Jan, L. Y., and Jan, Y. N. (1990) Similarity of the product of the Drosophila<br />

neurogenic gene big brain to transmembrane channel proteins. Nature 345, 163–167.<br />

30. Fortin, M. G., Morrison, N. A., and Verma, D. P. S. (1987) Nodulin-26, a peribacteroid<br />

membrane nodulin is expressed independently of the development of the peribacteroid<br />

compartment. Nucleic Acids Res. 15, 813–824.<br />

31. Feng, D.-F. and Doolittle, R. F. (1990) Progressive alignment and phylogenetic tree<br />

construction of protein sequences, in Methods in Enzymology, vol. 183: Molecular Evolution:<br />

Computer Analysis of Protein and Nucleic Acid Sequences (Doolittle, R. F., ed.),<br />

Academic, New York, pp. 375–387.


cDNA Libraries from Few Cells 499<br />

68<br />

cDNA Libraries from a Low Amount of Cells<br />

Philippe Ravassard, Christine Icard-Liepkalns, Jacques Mallet,<br />

and Jean Baptiste Dumas Milne Edwards<br />

1. Introduction<br />

1.1. Application of the SLIC Strategy to cDNA Libraries:<br />

General Considerations<br />

Conventional cDNA library construction often requires a minimum available amount<br />

of material (typically 1 or 2 µg of polyA + RNA). For complex organs, such as brain, or<br />

certain species, such as humans, as well as subsets of cell types, this condition is often<br />

difficult to fulfill. Amplification by polymerase chain reaction (PCR) can be used to<br />

circumvent this limitation because it is a powerful method to obtain working quantities<br />

of low-abundance DNAs. To effectively apply this method, known sequences need to<br />

be attached to the ends of the single-stranded cDNA (ss-cDNA). One at the 5′ end of<br />

the ss-cDNA is added during the priming of the synthesis; the other, at the 3′ end, is<br />

covalently attached by ligation using the SLIC strategy. With known DNA sequences<br />

attached to both ends of the synthetized cDNA, minute quantities can be amplified with<br />

sequence-specific primers to provide sufficient material to successfully generate and<br />

screen cDNA libraries. The overall scheme is illustrated in Fig. 1.<br />

Obviously, the goal in constructing such a library is to maintain the representation<br />

of every mRNA in the total RNA population, even low-abundant ones. Because the<br />

reverse transcription and the single-strand ligation do not modify the overall proportion<br />

of each molecule, only the PCR step may introduce an important bias in cDNA<br />

representation. Thus, for cDNA library construction using this methodology, it is<br />

important to optimize the synthesis of the sequence-tagged cDNA and to take steps<br />

to limit any amplification bias.<br />

1.2. Simultaneous PCR Amplification of a Complex DNA Mixture<br />

Generates an Important Size Bias<br />

The sequence-tagged cDNAs correspond to a large population of molecules with<br />

indentical ends. The only difference between these molecules is their relative size and<br />

sequence. Therefore, keeping the representativity of the cDNAs after PCR requires a<br />

constant amplification yield regardless of the size and sequence of the original cDNA<br />

population. This, unfortunately cannot be achieved by PCR. In fact, coamplification of<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

499


500 Ravassard et al.<br />

Fig. 1. Constructing cDNA libraries using the SLIC strategy. Reverse transcription is<br />

performed with a random primer RA3′ NV. RNAs are digested and primer is removed to avoid<br />

its ligation to the A5′ NV oligonucleotide. A5’ NV bears a phosphate group at its 5′ extremity<br />

to allow ligation to the 3′ end of the ss-cDNA. To avoid formation of concatemers A5′ NV<br />

bears an amino group at its 3′ end. Two rounds of nested PCR are performed to generate a<br />

ready-to-use cDNA library.<br />

short and long molecules with the same primers always leads to a selective amplification<br />

of the shorter ones, even though the longer ones were originally more abundant (1).<br />

Attempts to enlarge the average size of the PCR products have been made with the<br />

help of long-range PCR procedures, such as Pfu dilutions (2), Taq extender (Stratagene,<br />

La Jolla, CA), and Expand PCR system (Boehringer, Indianapolis, IN). The results<br />

were not significantly different if compared with the classic PCR techniques (data<br />

not shown).<br />

In conclusion, the size bias represents the major limitation in constructing PCR<br />

cDNA libraries. Thus, to avoid bias during the amplification step the average size of<br />

the cDNA library must be between 0.8 and 1 kbp, which is far below the usual average<br />

length of a ss-cDNA.


cDNA Libraries from Few Cells 501<br />

Fig. 2. Oligonucleotides used for 3′ anchored PCR. The three oligonucleotides A3′_1,2,3<br />

are designed from A3′ NV and RA3′ NV to be used in PCR experiments. Note that these three<br />

primers have to be used with A5′_1,2,3.<br />

1.3. ss-cDNA Synthesis<br />

The constraint caused by the size limitation of the ss-cDNA forbids the use of<br />

oligo-dT to prime the reverse transcription. If such a priming strategy is chosen, this<br />

will lead to a 3′-UTR-cDNA library. To obtain a cDNA library representative of the<br />

sequences of all messenger RNA, priming with random primer RA3′ NV must be<br />

performed.<br />

Experimentally, the ss-cDNA is synthesized with a random primer RA3′ NV (Fig. 2)<br />

and a radiolabled nucleotide. The average size is determined by alkaline gel electrophoresis<br />

and autoradiography. The incubation time with the reverse transcriptase<br />

is calibrated in order to generate an ss-cDNA whose average length is between 0.8<br />

and 1 kbp. With those conditions the representation of any mRNA will be optimal<br />

in the library.<br />

1.4. Amplification of the cDNA Library Based on the SLIC Strategy<br />

We have used this strategy to generate a cDNA library from newborn rat cervical<br />

superior ganglia (CSG). Total RNA was prepared from one single CSG, and polyA +<br />

RNA was prepared with oligo dT-coated magnetic beads (Dynabeads mRNA purification<br />

kit). Half of the material, which corresponds to polyA + RNA of about 5000 cells,<br />

was used to synthesize ss-cDNA primed with RA3′ NV. An incubation time of 30 min<br />

was optimum to generate an average size of 1 kbp. After removal of the primer, the<br />

ss-cDNA was ligated to A5′ NV (note that after each step the different primers are<br />

removed). These ss-cDNA were amplified by two rounds of nested PCR. To increase<br />

the specificity of the PCR reaction, we have used the touchdown PCR protocol (3). The<br />

primers used for the nested PCR were A5′_1 ∞ A3′_1 and A5′_2 ∞ A3′_2, respectively.<br />

One-twentieth of the reaction was cloned in a blunt-end vector, yielding 2 ∞ 10 5<br />

colonies. Analysis by direct PCR on colonies of 96 randomly chosen clones indicated<br />

that the average size of the library was about 900 bases. The striking result is that the


502 Ravassard et al.<br />

size dispersion around the mean is extremely low when compared with a conventional<br />

library. Finally, 5000 primary clones were screened with a TH oligonucleotide, yielding<br />

two positive TH clones (0.04%). Thus, no major distortion had been introduced in the<br />

abundance of TH clones, since TH represents 0.05% of CSG mRNA.<br />

1.5. Direct Screening of the PCR Library<br />

with <strong>Bio</strong>tinylated Oligonucleotides<br />

One of the major difficulties in making a cDNA library is the cloning of the<br />

double-stranded cDNA (ds-cDNA). We tested a direct screening protocol of uncloned<br />

ds-cDNA. After the second nested PCR, the amount of ds-cDNA was about 2 to 5 µg.<br />

We directly hybridized the denatured ds-cDNA with a biotinylated TH specific primer.<br />

After the hybridization reaction, probe-cDNA hybrids were separated from unhybridized<br />

DNA using streptavidin-coated magnetic beads (Dynabeads M-280). After various<br />

washing steps, the captured cDNA was amplified using the third nested PCR primers<br />

(A5′3 and A3′3) directly onto the beads. The PCR product was cloned, and more than<br />

85% of them were TH clones as analyzed by partial sequencing. <strong>Bio</strong>tinylated cDNA<br />

probes, instead of oligonucleotides, can also be used to screen a PCR library (4).<br />

1.6. Application of the SLIC Method to Subtractive Libraries<br />

Subtraction cloning strategies could be modified and certainly improved taking<br />

advantage of the generation of cDNA molecules exhibiting two defined extremities.<br />

Basically, tracer cDNA synthesized on mRNA from source A is hybridized to sequences<br />

of driver mRNA, which is isolated from a different but usually related source B. The<br />

tracer cDNAs that do not become hybridized with driver mRNA represent an enriched<br />

population of sequences expressed only in A cells. These are used for constructing an<br />

A-cell specific cDNA library.<br />

The SLIC strategy provides DNA molecules with two defined ends. This offers the<br />

opportunity to work on cDNA from populations A and B with two different sets of<br />

SLIC primers, A (A5′ NV and RA3′ NV) and B (B5′ NV and RB3′ NV; Fig. 3). During<br />

the PCR amplification of the tracer population B, a pair of biotinylated primers is used.<br />

Thus, this population can be captured and pure single-stranded molecules immobilized<br />

on magnetic beads. Then, after hybridization with the amplified A population, the<br />

unhybridized population corresponds to the A-cell specific sequences and can be used<br />

easily to generate substracted libraries or probes. This strategy gives for the first<br />

time the opportunity to realize such subtractive libraries with a very small amount<br />

of input material.<br />

1.7. Conclusion<br />

The SLIC method is a powerful and unique tool to synthesize cDNA and substractive<br />

libraries from a limited number of cells. It is unfortunately extremely difficult to<br />

generate full-length libraries. Nevertheless, cloning 5′ or 3′ ends of a cDNA is no<br />

longer a limiting step because anchored PCR can be easily performed. In this case, the<br />

same ss-cDNA that was used to generate the library can also be used as a matrix to<br />

isolate both ends of the incomplete clone.


cDNA Libraries from Few Cells 503<br />

Fig. 3. Alternative oligonucleotides used for subtractive cDNA library construction. For<br />

the B oligonucleotides the same modifications (5′ phosphate and 3′ amino) as the A primers<br />

have to be used.<br />

2. Materials<br />

2.1. Oligonucleotides<br />

1. PCR library primers: For the 5′ end of the ss-cDNA, use RA3′ NV and the related<br />

oligonucleotides (A3_1,2,3). For the 3′ end use A5′NV, A5′_1,2,3. All oligonucleotides<br />

must bear 5′ and 3′ hydroxyl group. Only A5′NV must have a phosphate group to its<br />

5′ end to allow ligation to the 3′ end of the ss-cDNA. To avoid selfligation of A5′ NV, its<br />

3′ end must be protected with an amino group. Those modifications are performed by any<br />

oligonucleotide suppliers. Apply the same rules when using B5′ NV and RB3′ NV. Note<br />

that only two related oligonucleotides are available for both ends.<br />

2. <strong>Bio</strong>tinylated screening primers: Ask your oligonucleotide supplier for 5′ biotinylated<br />

primers with a seven carbon spacer. Do not use an 11 carbon spacer because this will<br />

dramatically decrease the capture yield. If this primer is designed to hybridize in the middle<br />

of a molecule, add 4 to 6 random nucleotides at the 5′ end to facilitate the interaction with<br />

the streptavidin magnetic beads.<br />

2.2. RNA Extraction (see Note 1)<br />

1. Starting material: Fresh pelleted cells, store at –80°C on collection. Fresh pieces of organs,<br />

or tissue frozen in liquid nitrogen and stored at –80°C.<br />

2. PolytronR TP1200 (if organs are used).<br />

3. RNAZol reagent (BIOPROBE).<br />

4. DT40 (Pharmacia, [Piscataway, NJ]; #17-0270-01) Prepare a 5 mg/mL stock solution in<br />

ddH 2 O. Aliquot and store at –20°C.<br />

5. CHCl 3 .<br />

6. Isopropanol.<br />

7. 100% Ethanol.<br />

8. 70% Ethanol.


504 Ravassard et al.<br />

9. ddH 2 O.<br />

10. Dynabeads mRNA purification kit (Dynal, Lake Success, NY).<br />

2.3. Synthesis of the ss-cDNA<br />

1. Water bath at 70 and 42°C.<br />

2. Dry-ice powder and ice.<br />

3. 10× FSB buffer: 1 M Tris-HCl, pH 8.4 (at 42°C), 1.2 M KCl, 100 mM MgCl 2 . Aliquot<br />

and store at –20°C.<br />

4. DTT (100 mM ).<br />

5. Acetylated bovine serum albumin, 5 mg/mL (RNase-free; Life Technologies/BRL,<br />

Gaithersburg, MD).<br />

6. RNasin, 36 U/µL (Promega <strong>Bio</strong>tech, Madison, WI).<br />

7. dNTPs, 10 mM each (Use lithium-free dNTPs).<br />

8. Sodium pyrophosphate (20 mM; PPi).<br />

9. [α 32 P] dATP, 3000 Ci/mmol (Amersham, Arlington Heights, IL).<br />

10. AMV reverse transcriptase, 10 U/µL (Promega <strong>Bio</strong>tech, Madison, WI).<br />

11. RA3′ NV (50 ng/µL).<br />

2.4. Synthesis Yield Determination<br />

1. DE81 Whatman paper.<br />

2. Na 2 HPO 4 (0.5 M ).<br />

3. 100% Ethanol.<br />

4. Aqueous scintillation cocktail.<br />

2.5. Removal of Primers<br />

1. Prep-A-Gene DNA purification Kit (<strong>Bio</strong>-Rad [Richmond, CA]; #732-6010). The silica<br />

matrix used in this kit does not bind RNA or small DNA molecules (cut off around 100<br />

nucleotides) under oxidizing conditions.<br />

2. Water baths at 90 and 65°C.<br />

2.6. Ligation of the ss-cDNA to the Modified Oligonucleotide<br />

Modified oligonucleotide: 5′ phosphate and 3′ NH 2 A5′NV (or B5′NV).<br />

2.7. PCR Amplification<br />

PCR reagents, including primers.<br />

2.8. Direct Screening of the PCR Library<br />

with <strong>Bio</strong>tinylated Oligonucleotides<br />

1. <strong>Bio</strong>tinylated primer (see Subheading 2.1.2.).<br />

2. Streptavidin Dynabeads (10 mg/mL Dynabeads M-280 Streptavidin, DYNAL, Lake<br />

Success, NY).<br />

3. Magnetic concentrator (Dynal MPC).<br />

4. Rotating wheel.<br />

5. 20× SSPE solution: 200 mM NaH 2 PO 4 , pH 7.4, 3.6 M NaCl, 20 mM EDTA (5).<br />

6. 20× SSC: 300 mM sodium citrate, pH 7.0, 3 M NaCl (5).<br />

7. 50× Denhardt’s solution: 1% each of BSA, Ficoll 400, and PVP (5).<br />

8. 20% SDS.<br />

9. Sonicated salmon sperm DNA.


cDNA Libraries from Few Cells 505<br />

10. 10× PCR buffer (see Chapter 1).<br />

11. 42°C incubator.<br />

2.9. Blunt-End Cloning of PCR Products<br />

1. Unpurified PCR products.<br />

2. T4 DNA polymerase (4 U/µL, Amersham).<br />

3. Agarose gel (choose agarose concentration according to the PCR product length).<br />

4. QIAEX II purification kit (Qiagen, Inc., Chatsworth, CA).<br />

5. ATP (3 and 8 mM ).<br />

6. T4 polynucleotide kinase 5 U/µL (Amersham).<br />

7. 10× T4 polynucleotide kinase buffer (Amersham).<br />

8. Dephosphorylated SmaI pUC19 vector (Appligene).<br />

9. T4 DNA ligase, 4 U/µL.<br />

10. Electrocompetent XLI blue cells.<br />

11. Escherichia coli electroporation equipment.<br />

2.10. Direct PCR on Colonies<br />

1. Inoculating needles.<br />

2. PCR reagent, including primers. With pUC vectors, use M13 universal sequencing primer<br />

and M13 reverse sequence primer.<br />

3. Methods<br />

3.1. RNA Extraction<br />

All manipulations must be performed in an RNase-free environment and with PCR<br />

anticontamination material. In our hands, the best extraction yield for low amount of<br />

material is obtained with the RNAZol kit. When working with tissues, use a polytronR<br />

TP1200 to homogenize in the RNAzol solution.<br />

Follow the supplier’s instructions with two important modifications.<br />

1. Add 1 µL of DT40 (5 mg/mL) in the RNAZol solution prior to the homogenization step.<br />

This will increase the extraction yield.<br />

2. At the end, resuspend the pellet in 20 µL of ddH 2 O and store at –80°C.<br />

If the extraction of polyA + RNA is required, use the Dynabeads mRNA purification<br />

kit (Dynal); do not use the Dynabeads mRNA DIRECT kit. Elute from the magnetic<br />

beads with 20 µL of ddH 2 O instead of elution buffer and store at –80°C.<br />

3.2. Synthesis of the ss-cDNA (see Note 2)<br />

The final volume for the reverse transcription is 50 µL.<br />

3.2.1. AMV RTase Preincubation Mix<br />

1. Add, on ice, in a sequential manner the following reagents: 1.5 µL of H 2 O, 3 µL of 10× FSB,<br />

2.5 µL of 0.1 M DTT, 3 µL of 10 mM each dNTP (lithium free), 1 µL of 5 mg/mL BSA,<br />

5 µL [α 32 P] dATP (100 U/µL) (3000 Ci/mmol), 1 µL of RNasin (36 U/µL), 10 µL of<br />

20 mM PPi, and 1 µL of AMV RTase.<br />

2. Preincubate on ice for 30 min.


506 Ravassard et al.<br />

3.2.2. RNA Mix Preparation<br />

Prepare this mix during the preincubation of the AMV reverse transcriptase.<br />

1. Dilute the RNA in 17 µL of ddH 2 O (see Note 3).<br />

2. Add 1 µL of 50 ng/µL RA3′NV (or any anchored random primer).<br />

3. Add 2 µL of 10× FSB.<br />

4. Heat the tubes at 70°C for 15 min.<br />

5. Spin and freeze in dry-ice powder.<br />

6. Let it thaw on ice.<br />

3.2.3. Reverse Transcription<br />

1. Assemble the preincubation mix and the RNA mix and incubate at 42°C for 20 min<br />

to 1 h.<br />

2. To stop the reaction, add 1 µL of 0.5 M EDTA. To determine the optimal incubation time,<br />

perform five cDNA synthesis in 10 µL final reaction volume and stop the incubation every<br />

10 min after 20 min initial reaction time. Load 5 to 10 µL on an alkaline agarose gel (5)<br />

and measure the average length. The optimal reaction time will be the one that gives an<br />

average size of 0.8 to 1 kbp.<br />

3.3. Synthesis Yield Determination<br />

1. Take 1 µL of the cDNA and dilute it to 10 µL in ddH 2 O.<br />

2. Spot 5 µL of the dilution on two pieces of DE81 Whatman filter.<br />

3. Wash one filter only in 0.5 M Na 2 HPO 4 for 10 min.<br />

4. Repeat step 3 two more times and dry this filter in 100% ethanol.<br />

5. Dry both filters in air for 15 min.<br />

6. Add 10 mL of aqueous scintillation cocktail to the washed and unwashed filters.<br />

7. Count the 32 P activity. The washed filter activity corresponds to the incorporated activity<br />

(I), whereas the unwashed one corresponds to the total activity (T).<br />

8. The synthesized ss-cDNA mass (M) is given by the formula:<br />

M = (I / T) × (total dATP mass during reaction) × 4 (1)<br />

In the conditions used, M(ng) = (I / T) × 792. The overall yield should be between<br />

25 and 30% of the starting mass of RNAs.<br />

3.4. Removal of Primers<br />

Use Prep-A-Gene DNA purification kit. Follow the supplier’s instructions with the<br />

following important modifications.<br />

1. To remove primer after the ss-cDNA synthesis, heat for 5 min at 90 to 95°C to denature<br />

the RNADNA heteroduplexes. Add 150 µL of binding buffer, mix, then add 5 µL of<br />

resuspended matrix. Mix well. Incubate for 10 min at room temperature.<br />

2. To remove primer after ligation or PCR, do not perform this denaturation step.<br />

3. For purification do not use a starting volume smaller than 50 µL.<br />

4. Carefully remove all the wash buffer after the last wash. Traces of ethanol can be removed<br />

by drying the tubes for 3 min in a SpeedVac or equivalent rotary vacuum desiccator.<br />

5. Elute with 5 to 10 µL of ddH 2 O for 5 min at 65°C, then spin for 30 s and collect<br />

supernatant.<br />

6. Alternative procedures can be followed.


cDNA Libraries from Few Cells 507<br />

3.5. Ligation of the ss-cDNAs to the Modified Oligonucleotide<br />

(see Note 4)<br />

1. Remove primer before the PCR amplification step.<br />

3.6. PCR Amplification (see Notes 5 and 6)<br />

The general conditions used for both PCR amplifications are: 50 µL of reaction<br />

volume, hot start, and touchdown PCR.<br />

1. First PCR: Use A5′_1 and A3′_1 and half of the purified ligation mixture.<br />

2. Second PCR: Use A5′_2 and A3′_2 and one tenth of the purified first PCR.<br />

3. For both PCRs:<br />

a. Use as final concentration 200 µM of dNTPs, 0.8 µM of each primer, and 1.5 mM<br />

of MgCl 2 .<br />

b. Hot start: Add 0.5 µL of Taq DNA polymerase (5 U/µL) below the mineral oil when<br />

the reaction mixture reaches 80°C.<br />

c. Perform the following touch down PCR cycles: denaturation 93°C for 3 min; two cycles<br />

of 94°C, 30 s/70°C, 45 s/72°C, 1.5 min; two cycles of 94°C, 30 s/69°C, 45 s/72°C,<br />

1.5 min; two cycles of 94°C, 30 s/68°C, 45 s/72°C, 1.5 min; two cycles of 94°C,<br />

30 s/67°C, 45 s/72°C, 1.5 min; two cycles of 94°C, 30 s/66°C, 45 s/72°C, 1.5 min;<br />

25 cycles of 94°C, 30 s/65°C, 45 s/72°C, 1.5 min; and cool down to 4°C.<br />

d. Remove primer after each PCR.<br />

After the second nested PCR the amount of ds-cDNA is about 2 to 5 µg.<br />

3.7. Direct Sceening of the PCR Library<br />

with <strong>Bio</strong>tinylated Oligonucleotides<br />

3.7.1. Hybridization with the <strong>Bio</strong>tinylated Oligonucleotide (see Note 7)<br />

1. Use 500 ng of purified ds-cDNA. The volume should not exceed 8 µL.<br />

2. Add 2 µL of the 100 ng/µL biotinylated oligonucleotide.<br />

3. Adjust volume to 10 µL.<br />

4. Heat denature for 5 min at 95°C.<br />

5. Immediately add 100 µL of hybridization buffer: 5× SSPE, 5× Denhardt’s solution,<br />

1% SDS.<br />

6. Incubate overnight at 42°C.<br />

3.7.2. Separation of Probe-cDNA Hybrids<br />

After the hybridization reaction, probe-cDNA hybrids are separated from unhybridized<br />

DNA using Streptavidin-coated magnetic beads.<br />

1. Prehybridize Dynabeads with salmon sperm DNA by washing 20 µL of 10 mg/mL<br />

Dynabeads twice with 50 µL of hybridization buffer containing 250 µg/mL salmon sperm<br />

DNA. Incubate for 2 h at room temperature on a rotating wheel.<br />

2. Mix the Dynabeads with the probe-cDNA solution. Incubate 15 to 30 min at room<br />

temperature on a rotating wheel.<br />

3. The hybrids captured by the beads are washed twice with 1× SSC, 1% SDS, then twice<br />

with 0.1× SSC, 1% SDS. Washes are performed at 42°C for 20 min each.


508 Ravassard et al.<br />

4. Wash twice with 1× PCR buffer, 5% SDS, for 5 min at room temperature.<br />

5. Wash with 1× PCR buffer until SDS is completely removed. To do this, change the<br />

microtube after every wash.<br />

3.7.3. PCR Amplification of the Captured cDNA<br />

1. Transfer one fourth of the beads with the captured cDNA into a PCR tube. Make sure<br />

that all traces of SDS are removed.<br />

2. Perform PCR amplification with A5′_3 and A3′_3. Use the same protocol as described<br />

in Subheading 3.6.<br />

3.8. Blunt-End Cloning of PCR Products<br />

For blunt-end cloning, the 3′ overhanging extremities of the PCR product are<br />

removed with T4 DNA Polymerase (3′–5′ exonuclease activity). Oligonucleotides<br />

usually have 5′ hydroxyl ends. To allow ligation of the PCR product those extremities<br />

have to be phosphorylated by T4 Polynucleotide kinase (T4 PNK).<br />

1. At the end of the amplification reaction, add to the PCR mixture 1 µL of T4 DNA<br />

polymerase (4 U/µL). Incubate for 20 min at 16°C. Do not allow the temperature to<br />

rise above 16°C.<br />

2. Load the PCR product on a preparative agarose gel.<br />

3. Cut the desired bands and purify the DNA with the QIAEX II purification kit. Follow<br />

the supplier’s recommendations.<br />

4. Elute DNA from the silica matrix with 10 µL of ddH 2 O.<br />

5. Add 1.5 µL of 10× T4 PNK buffer, 1 µL of 3 mM ATP, 1.5 µL of ddH 2 O, and 1 µL of<br />

T4 PNK (5 U/µL).<br />

6. Incubate at 37°C for 30 min.<br />

7. Heat inactivate the enzyme at 75°C for 20 min.<br />

8. The PCR product is ready for ligation. Use the same buffer as the phosphorylation reaction,<br />

a dephosphorylated blunt-end vector (e.g., pUC19 SmaI), and a final concentration of<br />

ATP of 0.8 mM.<br />

9. Transform by electroporation and plate on the appropriate selection medium.<br />

3.9. Direct PCR on Colonies (see Note 8)<br />

1. Pick a colony with an inoculating needle.<br />

2. Touch the bottom of a 0.5-mL microtube (or a well in a microtiter plate) with the needle.<br />

3. Inoculate with the same needle, 3 mL of liquid bacterial growth medium in a 15-mL tube,<br />

and incubate at 37°C for 18 h.<br />

4. Repeat steps 1 to 3 for all the colonies to be analyzed.<br />

5. Prepare the PCR mixture on ice as follows. The volumes given are sufficient for one<br />

reaction: 2.5 µL of 10× PCR buffer, 2 µL of 2.5 mM dNTP, 2 µL of 50 mM MgCl 2 , 2 µL<br />

of primer 1 (50 ng/µL), 1 µL of primer 2 (50 ng/µL), 0.1 U of Taq DNA polymerase,<br />

and ddH 2 O to 25 µL.<br />

6. Distribute the PCR mixture into every tube on ice and add 100 µL of mineral oil.<br />

7. Place the tubes or the microtiter plate in the thermal cycler and run the following program:<br />

3 min denaturation at 93°C; 35 cycles of 94°C for 30 s, 55°C for 45 s, and 72°C for<br />

1 min/kb.<br />

8. Analyze the PCR products on an agarose gel.<br />

9. Prepare plasmid DNA of the positive clones from the cultures (prepared in step 3). Use<br />

this plasmid DNA for sequencing.


cDNA Libraries from Few Cells 509<br />

4. Notes<br />

1. All the material used in this manipulation must be very clean and at least sterilized. Wear<br />

gloves throughout the manipulation to avoid RNase contamination. The same precautions<br />

should be followed in the synthesis of the ss-cDNA. These precautions represent the lower<br />

level of protection against RNase and it is advisable to read ref. 6 carefully. Because<br />

PCR has to be performed later on this material, use anticontamination tips and aliquot<br />

every solution.<br />

2. During this manipulation, prepare controls that will be used in the ligation and PCR<br />

experiments. Prepare samples without AMV RTase and samples without RNA. This will<br />

lead to three different controls.<br />

3. We have successfully used as little as 10 ng of polyA + RNA. Do not exceed 1 µg of<br />

total RNA.<br />

4. Prepare nonligated samples composed of the same mixture as described in Subheading 3.5.<br />

without the T4 RNA ligase. Include each control of the cDNA synthesis.<br />

5. After removal of the primers, perform PCR amplification. Do not forget to include a PCR<br />

control without DNA for both amplifications.<br />

6. To analyze each PCR amplification, load 5 µL of (one tenth) of the PCR product on an<br />

agarose gel. After ethidium bromide staining, a signal could be observed after the first<br />

PCR but it is generally obtained after the second nested PCR.<br />

7. We have used degenerated primers to screen the PCR library. The amount of primer and<br />

Dynabeads should be at least five times above the amount described in Subheading 3.7.<br />

8. We have developped a direct PCR analysis to determine the sizes of inserted fragments<br />

and to rule out false positive recombinant clones.<br />

References<br />

1. Boularand, S., Darmon, M. C., and Mallet, J. (1995) The human tryptophan hydroxylase<br />

gene: An unusual complexity in the 5′ untranslated region. J. <strong>Bio</strong>l. Chem. 270, 3748–3756.<br />

2. Barnes, W. M. (1994) PCR amplification of up to 35-kb DNA with high fidelity and high<br />

yield from λ bacteriophage templates. Proc. Natl. Acad. Sci. USA 91, 2216–2220.<br />

3. Don, R. H., Cox, P. T., Wainwright, B. J., Baker, K., and Mattick, J. S. (1991) “Touch<br />

down” PCR to circumvent spurious priming during gene amplification. Nucleic Acids<br />

Res. 19, 4008.<br />

4. Abe, K. (1992) Rapid isolation of desired sequences from lone linker PCR amplified cDNA<br />

mixtures: Application to identification and recovery of expressed sequences in cloned<br />

genomic DNA. Mammalian Genome 2, 252–259.<br />

5. Sambrook, J., Fritsch, E. F., and Maniatis, T. (1989) Molecular Cloning: A Laboratory<br />

Manual, 2nd ed., Cold Spring Harbor Laboratory, Cold Spring Harbor, NY [SSPE, p. B.13,<br />

SSC, p. B.13, Denhart’s solution, p. B.15, Alkaline gel electrophoresis, p. B.23].<br />

6. Blumberg, D. D. (1987) Creating a ribonuclease free environment. Methods Enzymol.<br />

152, 20–24.


510 Ravassard et al.


Chimieric Junctions, Deletions, and Insertions 511<br />

69<br />

Creation of Chimeric Junctions, Deletions,<br />

and Insertions by PCR<br />

Genevieve Pont-Kingdon<br />

1. Introduction<br />

Recombinant polymerase chain reaction (PCR) (1) is the method of choice if one<br />

wants to modify a cloned DNA. It is a versatile technique that allows operations as<br />

different as creation of deletions, addition of small insertions, site-directed mutagenesis,<br />

and construction of chimeric molecules at any chosen location in the molecule of<br />

interest (see Note 1). This chapter describes in detail a simplification of the original<br />

recombinant PCR method. This fast and efficient method has been successful in fusing<br />

two different sequences with precision (2–4). It can also be used to create deletions<br />

or insert small fragments of DNA.<br />

The method (see Fig. 1) relies on a “chimeric primer” (C) and two outside primers<br />

(A and B). The final product can be obtained in one or two rounds of PCR. The figure<br />

illustrates the construction of a chimeric molecule in which two different templates<br />

are joined. The creation of a deletion, or the introduction of a small insertion within<br />

a given template, would follow the same pathway (see Note 1). In all cases, the new<br />

junction is designed in the chimeric primer; the 3′ half of the chimeric primer pairs<br />

with one of the templates (or one side of the deletion/insertion point), and its 5′ half has<br />

homology with the other template (or the other side of the deletion/insertion point).<br />

Both templates and the three primers (A–C) are placed in a reaction tube (step<br />

1) and one PCR is performed. During the first cycles, only the primers A and C can<br />

prime exponential amplification (step 2). This amplification reaction gives rise to an<br />

“intermediate fragment” (step 3), that can itself act as a primer. One of its 3′-ending<br />

strands anneals to the second template and is extended (step 4). This extension provides<br />

a template (step 5) for exponential amplification of the final product using the primers<br />

A and B (step 6). To limit the amplification of the intermediate fragment to the first<br />

cycles of PCR, the chimeric primer is used at a lower concentration than the two<br />

outside primers (3,4).<br />

2. Materials<br />

1. Primers (see Subheading 3.1.).<br />

2. DNA templates, linearized at a site outside the region to be amplified.<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

511


512 Pont-Kingdon<br />

Fig. 1. Construction of a chimeric product. See text for explanation of steps 1–6. The dsDNA<br />

templates are the plain and dotted double lines. The outside primers are short, plain (A), or<br />

dotted (B) single lines. The drawing of the chimeric primer (half plain, half dotted) reflects<br />

its homologies to the templates. Half arrowheads indicate 3′ ends. Closed arrows indicate<br />

amplification steps (2 and 6).<br />

3. GeneAmp PCR Core Reagent Kit (Perkin–Elmer, Foster City, CA) containing: AmpliTaq<br />

DNA polymerase (5 U/µL), Gene Amp dNTPs (10 mM solutions of dATP, dGTP, dTTP,<br />

and dCTP), GeneAMP 10× PCR buffer II (100 mM Tris-HCl, pH 8.3; 500 mM KCl),<br />

and 25 mM MgCl 2 solution.<br />

4. Mineral oil.<br />

5. Restriction endonucleases and buffers.<br />

6. Agarose gel electrophoresis reagents and equipment.<br />

7. Phenol:CHCl 3 isoamyl alcohol (25241, vvv).<br />

8. Ammonium acetate (7.5 M ).<br />

9. 100% Ethanol.<br />

10. 70% Ethanol.<br />

3. Methods<br />

3.1. Design of Chimeric Primer<br />

The chimeric primer is crucial because it contains the new junction and because<br />

sequences on each side of the junction serve as primers in different steps (i.e., 2 and 4<br />

in Fig. 1) of the reaction. To allow priming by the nucleotides found on each side of<br />

the junction, the new junction should be placed in the middle of an oligonucleotide<br />

of sufficient length. Otherwise, the design of a chimeric primer should follow the<br />

classical rules of primer-design (5). Our chimeric primer was a 34-mer, with 17 bases<br />

homologous to one template and 17 bases homologous to the other. Some slightly<br />

longer chimeric primers (36- and 40-mer) have been used (3,4). See Note 2 for more<br />

<strong>info</strong>rmation on the design of the chimeric primer.


Chimieric Junctions, Deletions, and Insertions 513<br />

3.2. Design of Outside Primers<br />

Fewer constraints apply to the design of the two outside primers, and therefore<br />

following the general rules of primer design should be adequate. In principle, the<br />

outside primers can be kilobase pairs away (within the limit of PCR feasibility) from<br />

the new junction. However, because their location determines the size of the product<br />

that will be cloned, the choice of their position is important. For several reasons, it<br />

is advantageous to plan the cloning of a fragment of few hundred nucleotides instead<br />

of a longer one: First, the smaller the fragment is, the less chance there is to find<br />

PCR-induced mutations in the final clone. Second, a smaller piece of DNA has to be<br />

sequenced to verify the integrity of the newly cloned DNA. This can be achieved by<br />

using primers that anneal close to the junction and have “built in” restriction sites<br />

at their 5′ ends. Another approach is to choose primers that anneal further from the<br />

junction and clone with restriction sites that closely surround the chimeric junction in<br />

the final product. The choice between these two possibilities depends on the cloning<br />

strategy and the availability of cloning sites in the amplification product. If the chimeric<br />

fragment has to be cloned into a new vector, restriction sites unique in both vector and<br />

chimeric fragment can be engineered at the 5′ ends of the two outside primers. If the<br />

chimeric junction has to be replaced by cloning in one of the original templates, restriction<br />

sites that exist in this template and in the chimeric junction have to be used.<br />

The size of the outside primers can be different than the size of the chimeric primer.<br />

We have been successful with a 34-mer chimeric primer, and two outside primers of 23<br />

and 20 nucleotides, respectively. If the size of the two outside primers is very different<br />

than half the size of the chimeric primer, series of cycles with different annealing<br />

temperatures can be performed (see Note 3).<br />

3.3. Procedure<br />

1. Assemble the components in one 50-µL reaction containing: both templates (10 fmol<br />

each), both outside primers (25 pmol each), chimeric primer (1 pmol), 50 µM each dNTP<br />

(see Note 4), 1× AmpliTaq DNA polymerase buffer, 1.5 mM MgCl 2 (see Note 5), and 2.5<br />

U of AmpliTaq DNA polymerase.<br />

2. Top PCR mix with 50 µL of mineral oil.<br />

3. Perform PCR as follows (see Note 3):<br />

a. Three to five initial cycles to allow initiation and amplification (steps 2–4 in Fig. 1)<br />

using the chimeric primer: 95°C for 30 s, T1 (see Note 3) for 30 s, and 72°C for 1 min<br />

for each kilobase of intermediate fragment.<br />

b. Twenty to thirty cycles to amplify the final chimeric product: 95°C for 30 s, T2 (see<br />

Note 3) for 30 s and 72°C for 1 min for each kilobase of chimeric product.<br />

4. Directly analyze 10 µL of the amplified DNA by standard procedures (restriction enzyme<br />

digests and electrophoresis, see Note 6).<br />

5. To clone the new junction, clean the PCR product by extraction with an equal volume of<br />

phenolchloroformisoamyl alcohol (25241).<br />

6. Transfer the aqueous phase to a sterile microcentrifuge tube. Remove the excess dNTP (it<br />

can inhibit T4 ligase) by ethanol precipitation. Add 1/2 vol of 7.5 M ammonium acetate<br />

and 2.5 vol of 100% ethanol.<br />

7. Mix well and incubate at room temperature for 10 min.<br />

8. Spin in a microfuge at maximum speed for 5 min.


514 Pont-Kingdon<br />

9. Invert tube and allow to drain.<br />

10. Wash pellet with 70% ethanol.<br />

11. Resuspend the DNA in 20–50 µL of TE. Quantify the DNA by UV spectrophotometry<br />

and use for cloning.<br />

4. Notes<br />

1. The applications of this method are diverse; it allows the creation of:<br />

a. Chimeric molecules: This method is well suited to cases in which the two molecules<br />

to be joined are unrelated. In the case where the two templates are fused in a region of<br />

homology, another PCR technique (6) might be preferred.<br />

b. Deletions: A set of different deletions can be easily obtained from the same template by<br />

using a set of different chimeric primers and only one set of outside primers.<br />

c. Insertions: The size of potential insertion using this technique is limited to the size of<br />

the chimeric primer. A restriction site sequence, sandwiched into a chimeric primer,<br />

could be introduced at will into any DNA.<br />

2. As in the Megaprimer method (7), the technique described here uses a PCR product as<br />

a primer. It has been mentioned for the Megaprimer method (8) that mutations can be<br />

found in the final product because of the tendency of Taq polymerase to add nontemplated<br />

nucleotides at the 3′ end of newly synthesized DNA strands. The frequency of these<br />

mutations should be low since 3′-ending DNA strands carrying nontemplated nucleotides<br />

should not prime well for the synthesis of the final product. Although we did not observe<br />

such mutations in the two final clones that we obtained and sequenced, this phenomenon<br />

could apply here, and we encourage the reader to refer to Note 1 in ref. 8 for a complete<br />

discussion.<br />

3. In this technique, each half of the chimeric primer must anneal with its target. We have<br />

limited the size of our chimeric primer to 34 nucleotides, giving us 17 nucleotides for<br />

each half. Because our outside primers are 20 and 23 nucleotides in length, we felt that it<br />

was necessary to plan a first set of cycles with a lower annealing temperature to allow the<br />

stable annealing of each half of the chimeric primer in the steps 2 and 4. This precaution is<br />

not necessary if all the sequences with a “priming” role (each half of the chimeric primer<br />

and both outside primers) are close in length and in G+C content.<br />

The temperature T1 is the approximate “annealing temperature” ([number of G + C ×<br />

4°C] + [number of A + T × 2°C] – 10°C) of the less stable half of the chimeric primer.<br />

The temperature T2 is the approximate annealing temperature of the less stable of the<br />

two outside primers.<br />

4. To limit the number of PCR-induced mutations in the final chimeric product, a low<br />

concentration of each dNTP (50 µM) is used.<br />

5. The optimal MgCl 2 concentration can vary among different pairs of primers. It is wise to<br />

define the best MgCl 2 concentration in a test experiment (9). In fact, it is possible that the<br />

two consecutive PCRs that occur in the tube have incompatible requirements for MgCl 2 .<br />

If this is the case, the chimeric product can be obtained by a two-step method (2). In this<br />

alternative, the intermediate fragment obtained during the first reaction is extracted with<br />

phenol:chloroform:isoamyl alcohol and purified from excess primers by precipitation with<br />

isopropanol from 2 M NH 4 OAc. The purified DNA is then used as a primer in a second<br />

reaction that contains the second template and the two outside primers.<br />

6. It is often stated that the oil that tops the PCR has to be removed to properly load the<br />

DNA into the well of an electrophoresis gel. We found that this step is not needed if the<br />

pipet tip is cleaned with tissue paper (Kimwipe) just after it has been filled with a DNA<br />

sample. In fact we found that the oil left in the tube allows for longer conservation of<br />

the sample at 4°C.


Chimieric Junctions, Deletions, and Insertions 515<br />

Acknowledgment<br />

I want to thank Janet E. Lindsley and Dana Carroll for helpful comments.<br />

References<br />

1. Higuchi, R. (1989) Using PCR to engineer DNA, in PCR Technology: Principles and<br />

Applications for DNA Amplification, Stockton, New York, pp. 61–70.<br />

2. Pont-Kingdon, G. (1994) Construction of chimeric molecules by a two-step recombinant<br />

PCR method. <strong>Bio</strong>Techniques 16, 1010–1011.<br />

3. Cao, Y. (1990) Direct cloning of a chimeric gene fused by the polymerase chain reaction.<br />

Technique 2, 109–111.<br />

4. Yon, J. and Fried, M. (1989) Precise gene fusion by PCR. Nucleic Acids Res. 17, 4895.<br />

5. Sharrocks, A. D. (1994) The design of primers for PCR, in PCR Technology: Current<br />

Innovations (Griffin, G. G. and Griffin, A. M., eds.), CRC, Boca Raton, FL, pp. 5–11.<br />

6. Klug, J., Wolf, M., and Beato, M. (1991) Creating chimeric molecules by PCR directed<br />

homologous DNA recombination. Nucleic Acids Res. 19, 2793.<br />

7. Sarkar, G. and Sommer, S. S. (1990) The “megaprimer” method of site-directed mutagenesis.<br />

<strong>Bio</strong>techniques 8, 404– 407.<br />

8. Barik, S. (1993) Site-directed mutagenesis by double polymerase chain reaction, in Methods<br />

in Molecular <strong>Bio</strong>logy, vol. 15: PCR Protocols: Current Methods and Applications (White,<br />

B. A., ed.), Humana, Totowa, NJ, pp. 277–286.<br />

9. Saiki, R. K. (1989) The design and optimization of the PCR, in PCR Technology: Principles<br />

and Applications for DNA Amplification (Erlich, H. A., ed.), Stockton, New York,<br />

pp. 7–16.


516 Pont-Kingdon


Recombinant PCR 517<br />

70<br />

Recombination and Site-Directed Mutagenesis<br />

Using Recombination PCR<br />

Douglas H. Jones and Stanley C. Winistorfer<br />

1. Introduction<br />

The polymerase chain reaction (PCR) (1) provides a rapid means for the recombination<br />

and site-directed mutagenesis of DNA (2). DNA modification can occur during<br />

PCR because the primers are incorporated into the ends of the PCR product. The<br />

simplest PCR-based method for site-directed mutagenesis and DNA recombination<br />

is recombination PCR.<br />

Recombination PCR uses in vivo recombination in Escherichia coli to generate sitedirected<br />

mutants and recombinant constructs (3,4). In the recombination PCR method,<br />

PCR adds homologous ends to DNA. These homologous ends mediate recombination<br />

between these linear PCR products in E. coli, resulting in the formation of DNA joints<br />

in vivo. If two PCR products have homologous ends that can recombine to form a circle,<br />

and if this circle constitutes a selectable plasmid, E. coli can be readily transformed<br />

by the linear PCR products. Recombination PCR has almost no steps apart from PCR<br />

amplification and transformation of E. coli, and this method works well in Rec A minus<br />

E. coli strains used routinely in cloning. Since the introduction of this method in 1991,<br />

it has been used by numerous investigators (5–9). One example of DNA recombination<br />

using recombination PCR is illustrated in Fig. 1, which shows a protocol for amplifying<br />

a portion of a donor plasmid and inserting it in a recipient plasmid at a defined position<br />

and orientation. The donor plasmid is shown on the left side and the recipient plasmid is<br />

on the right side. The steps corresponding to this figure are briefly outlined below:<br />

1. The DNA segment that is to be inserted into the recipient plasmid is amplified from<br />

the donor plasmid using primers 1 and 2. In a separate PCR amplification, the recipient<br />

plasmid is amplified with primers 3 and 4. The 5′ regions of primers 1 and 2 that do not<br />

anneal to the donor plasmid are complementary to primers 4 and 3, respectively (or 3<br />

and 4, depending on the orientation of the insert desired in the recombinant construct).<br />

Frequently, a plasmid template can be linearized outside the region to be amplified by<br />

restriction endonuclease digestion before PCR amplification. When this can be done,<br />

the PCR product does not need to be purified, because linearized plasmids transform E. coli<br />

inefficiently. If a plasmid template cannot be linearized by restriction endonuclease digestion<br />

outside the region to be amplified prior to PCR amplification, the PCR product must<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

517


518 Jones and Winistorfer<br />

Fig. 1. Diagram illustrating DNA recombination using recombination PCR. The primers are<br />

numbered hemiarrows. The insert is the cross-hatched region. Smooth circles represent the DNA<br />

strands of the donor plasmid. Circles with wavy and jagged portions represent DNA strands of<br />

the recipient plasmid. Reprinted by permission from <strong>Bio</strong>techniques 10, 62–66.<br />

be removed from the plasmid before transformation to prevent background transformants<br />

arising from the supercoiled plasmid template. PCR product purification is accomplished<br />

either by agarose gel purification followed by glass bead extraction or by adding the<br />

restriction endonuclease DpnI to the PCR mixture. DpnI is a restriction endonuclease that<br />

digests methylated GATC sites. These sites are methylated in the plasmid by strains of<br />

E. coli used routinely in cloning (by dam methylase), but are not methylated in the PCR<br />

products, permitting DpnI to digest the plasmid without cutting the PCR product (10).<br />

2. The two PCR products are combined and used to transform MAX efficiency competent<br />

E. coli (BRL, Life Technologies, Gaithersburg, MD). If each plasmid template is restriction<br />

endonuclease digested outside the region to be amplified prior to PCR amplification,<br />

the two crude PCR products can simply be combined, and the resulting mixture used<br />

to transform E. coli.<br />

In a simple variation of this recombination PCR strategy, the inserted segment can be an<br />

unmodified PCR product. In that case, primers 3 and 4 have 5′ ends that are homologous<br />

to the ends of the PCR fragment to be inserted, and the recipient plasmid is linearized<br />

by restriction endonuclease digestion prior to PCR amplification. We routinely use this<br />

approach to clone any PCR product (4).


Recombinant PCR 519<br />

In recombination PCR, the sum goal of the two amplifications is to yield two PCR<br />

products where each end of one product is homologous to a distinct end of the other<br />

PCR product. Because the amplifying primer sequences are incorporated into the ends of<br />

a PCR product, so long as primers 1 and 2 contain regions that are complementary to<br />

regions of primers 3 and 4 (or 4 and 3), the PCR products will contain ends that are<br />

homologous to each other, and these primer-determined DNA ends do not need to be<br />

determined by the original donor or recipient templates. The only requirement of this<br />

recombination PCR strategy is that primers 1 and 2 must have regions of complementarity<br />

to primers 3 and 4. Therefore, recombination PCR can be used not only to generate<br />

recombinant constructs, such as gene chimeras, but also for the site-directed mutagenesis<br />

of two distal sites concurrently (Fig. 2) or for the rapid site-directed mutagenesis of single<br />

sites (Fig. 3) (11). In the point mutagenesis protocol illustrated in Fig. 3, the plasmid is<br />

linearized by restriction endonuclease digestion before each PCR amplification. In each<br />

of the two amplifications, the mutating primers (primers 1 and 3) mutate the identical<br />

base pair so that the mutated ends of each product are homologous to each other and<br />

the nonmutating primers (primers 2 and 4) are also designed to produce ends that are<br />

homologous to each other. Both unpurified PCR products are combined to transform<br />

E. coli, generating clones with the mutation of interest.<br />

2. Materials<br />

1. Taq DNA polymerase (AmpliTaq 5 U/mL; Perkin–Elmer, Norwalk CT) (see Note1).<br />

2. 10× PCR buffer II: 500 mM KCl, 100 mM Tris-HCl, pH 8.3.<br />

3. MgCl 2 solution (25 mM).<br />

4. Stocks (10 mM ) of each dATP, dCTP, dTTP, and dGTP, neutralized to pH 7.0 with NaOH.<br />

5. Restriction endonucleases (New England <strong>Bio</strong>Labs, Beverly, MA).<br />

6. PCR primers. In Fig. 1, PCR amplification with primers 1 and 2 results in a product with<br />

24 to 30 bp of homology with the products of primers 3 and 4. For PCR primers that<br />

introduce mutations, see Note 2.<br />

7. Agarose.<br />

8. Ethidium bromide.<br />

9. TAE buffer: 40 mM Tris-acetate, 2 mM EDTA, pH 8.5 (12).<br />

10. Geneclean (<strong>Bio</strong> 101, La Jolla, CA).<br />

11. TE buffer: 10 mM Tris-HCl, pH 8.0, 1 mM EDTA.<br />

12. MAX Efficiency DH5α Competent E. coli (BRL, Life Technologies). Once a tube is<br />

thawed it should not be reused (see Note 3).<br />

13. SOC Media (13).<br />

14. LB plates with 100 µg/mL ampicillin (14).<br />

15. Luria-Bertani medium (LB broth) (15).<br />

3. Methods<br />

1. Linearize the plasmid template by restriction endonuclease digestion outside the region to<br />

be amplified, if possible. The plasmid digest does not need to be purified prior to its use<br />

as a PCR template (see Notes 4 and 5).<br />

2. Assemble a PCR in a total volume of 50 µL containing the following: 2 ng of plasmid<br />

template, 25 pmol of each primer, 200 µM each dNTP, 1X PCR buffer, 2.5 mM MgCl 2 ,<br />

and 1.25 U DNA Taq DNA polymerase. (see Note 6).<br />

3. Perform PCR amplification using the following parameters (see Note 6): 94°C for 1 min<br />

(initial denaturation), 94°C for 30 s (denaturation), 50°C for 30 s (anneal), 72°C for<br />

1 min/kb of PCR product (extension), 14–20 amplification cycles, and 72°C for 7 min<br />

(final extension step).


520 Jones and Winistorfer<br />

Fig. 2. Diagram illustrating site-directed mutagenesis of two distal sites using recombination<br />

PCR. The primers are numbered hemiarrows. Asterisks designate the mutagenesis sites. There<br />

is no purification of the PCR products. Notches designate point mismatches in the primers and<br />

resulting mutations in the PCR products. Reprinted by permission from Technique 2, 273–278.<br />

4. Visualize the PCR product on an agarose minigel. If 5 µL of the PCR product can be clearly<br />

seen following ethidium bromide staining (>15 ng/5 µL), there is enough product.<br />

5. Withdraw 2.5 µL from each PCR tube (typically 10–60 ng/2.5 µL) and then combine the<br />

two samples. If the PCR template is linearized by restriction endonuclease digestion outside<br />

the region to be amplified, no purification of PCR products is necessary (see Note 5).<br />

6. Transform MAX efficiency competent E. coli (BRL) with the 5-µL sample containing the two<br />

PCR products. Maintaining an even molar ratio of one product to another is not necessary.<br />

Transformation is carried out following the manufacturer’s instructions with the following<br />

modifications:


Recombinant PCR 521<br />

Fig. 3. Diagram illustrating site-directed mutagenesis of a single site using recombination<br />

PCR with 4 primers. The primers are numbered hemiarrows. The asterisk designates the<br />

mutagenesis site. Primer 2 is complementary to primer 4. Restriction endonuclease sites A and<br />

B bracket the insert. Notches designate point mismatches in the primers and resulting mutations<br />

in the PCR products. There is no purification of the PCR products. For each additional single<br />

site-directed mutagenesis reaction, only a new primer 1 and 3 need to be synthesized, and the<br />

same cut templates can be used. Reprinted by permission from Technique 2, 273–278.<br />

a. Use 50 µL of E. coli for each sample transformed, because this is effective and less<br />

expensive than the 100 µL recommended.<br />

b. After incubation at 37°C in a shaker for 1 h, plate the entire sample onto an LB plate<br />

containing 100 µg/mL ampicillin.<br />

c. Once an aliquot of bacteria is thawed, do not use it again.<br />

Then, set up the following transformations: Plate A: 2.5 µL of PCR 1 + 2.5 µL of PCR 2;<br />

Plate B: 2.5 µL of PCR 1 + 2.5 µL of TE; Plate C: 2.5 µL of PCR 2 + 2.5 µL of TE; Plate<br />

D: 0.5 ng of a supercoiled template in 5 µL of TE; and Plate E: 5 µL of TE.


522 Jones and Winistorfer<br />

The yield of colonies from plate A is >2 times that from plate B + C, confirming a high<br />

percentage of recombinants in plate A. Plate D is a transformation control, and should<br />

yield a thick lawn of colonies. Plate E is an antibiotic control, and should yield no colonies<br />

since the bacteria that have not been transformed are sensitive to ampicillin. Only 25 µL<br />

of cells are used for the control plates D and plate E, so that only one BRL tube, which<br />

contains 200 µL of bacteria, needs to be used.<br />

7. Screen plasmids by placing individual colonies in 2 mL of LB broth containing 100 µg/mL<br />

of ampicillin. Grow the colonies at 37°C for 6–24 h.<br />

8. Screen the plasmids by removing 2 µL of the LB broth, place it directly in a PCR tube,<br />

and amplify for 25 cycles (see steps 2–4) using primers that flank the mutated site or<br />

insert (e.g., M13 primers) (16).<br />

9. In a mutagenesis protocol, a base pair can be mutated to either create or remove a restriction<br />

endonuclease site. In particular, the degenerate amino acid code allows one to create<br />

or eliminate a restriction endonuclease recognition site without altering the amino acid<br />

encoded at that site. Screen for the mutation by adding 3 U of the restriction endonuclease<br />

and 1 µL of the appropriate 10× restriction buffer directly to 5 µL of the unpurified PCR<br />

product in a total volume of 10 µL.<br />

10. Assess cutting by minigel analysis. Typically, 50 to 100% of the clones contain the<br />

recombinant of interest. Then, purify the plasmid and sequence the mutated region (see<br />

Note 7).<br />

4. Notes<br />

1. Other investigators have used Pfu DNA polymerase instead of Taq DNA polymerase<br />

in recombination PCR (6). Pfu DNA polymerase has better fidelity than Taq DNA<br />

polymerase (17).<br />

2. The primers that introduce point mutations (primers 1–4 in Fig. 2 and primers 1 and 3 in<br />

Fig. 3) are designed to generate 15to 45 bp of homology between each end of one PCR<br />

product relative to the other PCR product. In all recombination PCR protocols, 24 bp of<br />

homology works very well, and alterations that generate long regions of homology do<br />

not work noticeably better. Decreasing the length of homology from 25 to 12 bp in an<br />

early protocol did decrease the transformation efficiency four- to five-fold. Single point<br />

mismatches lie no closer than six nucleotides from the 3′ end of a primer and are frequently<br />

placed toward the middle. Placing point mutations near the 5′ end of each mutating primer<br />

will generate two PCR products whose mutated ends have >24-bp of homology. Multiple<br />

point mismatches should be placed in the middle or toward the 5′ end of a primer, with<br />

primer lengths long enough to create 24-bp of homology between the mutated ends of the<br />

two PCR products. Primers that are nonmutating are generally 24 to 30 nucleotides long.<br />

These nonmutating primers can be designed to anneal to the β-lactamase gene so that they<br />

can be used with a variety of different plasmids. Frequently, the mutating and nonmutating<br />

primers are designed to be perfect complements to each other.<br />

For site-directed mutagenesis, since unique restriction endonuclease recognition sites<br />

almost always bracket the insert, the same linearized templates can be used for the<br />

mutagenesis of any single site in the insert. Primers 2 and 4 are nonmutating (see Fig. 3),<br />

and are conserved for each new site targeted for mutagenesis, so that only two new<br />

primers need to be generated for each site targeted for mutagenesis (via primers 1 and 3).<br />

Furthermore, only approx one half of the length of the entire template needs to be amplified<br />

in each of the two PCR amplifications, facilitating the mutagenesis of large constructs and<br />

permitting considerable flexibility in the primer design and sequence. Recombination PCR<br />

has been used to mutate constructs up to 7.1 kb (18).


Recombinant PCR 523<br />

3. Because the transformation efficiency is low, highly competent bacteria (transfection<br />

efficiency >1 × 10 9 /µg of monomer pUC19) should be used. Using restriction endonuclease<br />

digested templates, the transformation efficiency is about 10 colonies with the mutation/ng<br />

total DNA used to transform E. coli.<br />

4. After 14 amplification cycles, the PCR product yield is much higher when using a linear<br />

template than when using a supercoiled template.<br />

5. If a plasmid template cannot be linearized outside the region to be amplified before PCR<br />

amplification, the PCR product must be removed from the supercoiled plasmid template.<br />

This can be accomplished either by agarose gel electrophoresis and extraction using<br />

GeneClean or by digestion with the restriction endonuclease DpnI. When agarose gel<br />

resolution and GeneClean extraction are used, the entire PCR product should be gel<br />

purified and reconstituted in 25 to 30 µL of TE, and 2.5 µL is combined with the<br />

2.5 µL of the other PCR product before transformation. If DpnI is used, add 20 U of DpnI<br />

directly to 25 µL of the PCR sample using the recommended 10× DpnI buffer in a final<br />

total volume of 30 µL, and incubate the mixture at 37°C for 1 h. No further purification<br />

of the PCR product is necessary.<br />

6. The exact buffer components and conditions for PCR vary with different primers and<br />

template.<br />

7. There is always the possibility of a sequence error in a single clone after PCR amplification.<br />

The altered region should be sequenced, and one may choose to clone a restriction fragment<br />

containing the mutated or recombined region of interest into a construct that has not<br />

undergone PCR amplification.<br />

Acknowledgments<br />

This work was supported by the Roy J. Carver Charitable Trust, the University of<br />

Iowa through funds generated by the Childrens Miracle Network Telethon, and by<br />

National Institutes of Health grant R01 HG00569. We thank Jim Hartley for suggesting<br />

use of the restriction endonuclease DpnI to remove supercoiled template from the<br />

PCR mixture.<br />

References<br />

1. Mullis, K., Faloona, F., Scharf, S., Saiki, R., Horn, G., and Erlich, H. (1986) Specific<br />

enzymatic amplification of DNA in vitro: the polymerase chain reaction, in Cold Spring<br />

Harbor Symposia on Quantitative <strong>Bio</strong>logy, vol. LI, Cold Spring Harbor Laboratory, Cold<br />

Spring Harbor, NY, pp. 263–273.<br />

2. White, B. (1993) Methods in Molecular <strong>Bio</strong>logy, vol. 15, PCR Protocols: Current Methods<br />

and Applications, Humana, Totowa, NJ.<br />

3. Jones, D. H. and Howard, B. H. (1991) A rapid method for recombination and site-specific<br />

mutagenesis by placing homologous ends on DNA using polymerase chain reaction.<br />

<strong>Bio</strong>techniques 10, 62–66.<br />

4. Jones, D. H. (1994) PCR mutagenesis and recombination in vivo. PCR Methods Appl.<br />

3, S141–S148.<br />

5. Coco, W. M., Rothmel, R. K., Henikoff, S., and Chakrabarty, A. M. (1993) Nucleotide<br />

sequence and initial functional characterization of the clcR gene encoding a LysR family<br />

activator of the clcABD chlorocatechol operon in Pseudomonas putida. J. Bacteriol. 175,<br />

417– 427.<br />

6. Fridovich-Keil, J. L. and Jinks-Robertson, S. (1993) A yeast expression system for human<br />

galactose-1-phosphate uridylyltransferase. Proc. Natl. Acad. Sci. USA 90, 398– 402.


524 Jones and Winistorfer<br />

7. Goulden, M. G., Kohm, B. A., Santa Cruz, S., Kavanagh, T. A., and Baulcombe, D. C.<br />

(1993) A feature of the coat protein of potato virus X affects both induced virus resistance<br />

in potato and viral fitness. Virology 197, 293–302.<br />

8. Gibbs, J. S., Regier, D. A., and Desrosiers, R. C. (1994) Construction and in vitro properties<br />

of HIV-1 mutants with deletions in “nonessential” genes. AIDS Res. Hum. Retrovir. 10,<br />

343–350.<br />

9. Singh, K. K., Small, G. M., and Lewin, A. S. (1992) Alternative topogenic signals in peroxisomal<br />

citrate synthase of Saccharomyces cerevisiae. Mol. Cell. <strong>Bio</strong>l. 12, 5593–5599.<br />

10. Weiner, M. P., Costa, G. L., Schoettlin, W., Cline, J., Mathur, E., and Bauer, J. C. (1994)<br />

Site-directed mutagenesis of double-stranded DNA by the polymerase chain reaction.<br />

Gene 151, 119–123.<br />

11. Jones, D. H. and Winistorfer, S. C. (1992) Recombinant circle PCR and recombination<br />

PCR for site-specific mutagenesis without PCR product purification. <strong>Bio</strong>techniques 12,<br />

528–534.<br />

12. Sambrook, J., Fritsch, E. F., and Maniatis, T. (1989) Molecular Cloning: A Laboratory<br />

Manual, 2nd ed., Cold Spring Harbor Laboratory, Cold Spring Harbor, NY, p. B23.<br />

13. Sambrook, J., Fritsch, E. F., and Maniatis, T. (1989) Molecular Cloning: A Laboratory<br />

Manual, 2nd ed., Cold Spring Harbor Laboratory, Cold Spring Harbor, NY, p. A2.<br />

14. Sambrook, J., Fritsch, E. F., and Maniatis, T. (1989) Molecular Cloning: A Laboratory<br />

Manual, 2nd ed., Cold Spring Harbor Laboratory, Cold Spring Harbor, NY, p. A4.<br />

15. Sambrook, J., Fritsch, E. F., and Maniatis, T. (1989) Molecular Cloning: A Laboratory<br />

Manual, 2nd ed., Cold Spring Harbor Laboratory, Cold Spring Harbor, NY, p. A1.<br />

16. Liang, W. and <strong>John</strong>son, J. P. (1988) Rapid plasmid insert amplification with polymerase<br />

chain reaction. Nucleic Acids Res. 16, 3579.<br />

17. Lundberg, K. S., Shoemaker, D. D., Adams, M. W. W., Short, J. M., Sorge, J. A., and<br />

Mathur, E. J. (1991) High-fidelity amplification using a thermostable DNA polymerase<br />

isolated from Pyrococcus furiosus. Gene 108, 1–6.<br />

18. Yao, Z., Jones, D. H., and Grose, C. (1992) Site-directed mutagenesis of herpesvirus<br />

glycoprotein phosphorylation sites by recombination polymerase chain reaction. PCR<br />

Methods Appl. 1, 205–207.


Mutagenesis by Megaprimer PCR 525<br />

71<br />

Megaprimer PCR<br />

Application in Mutagenesis and Gene Fusion<br />

Emily Burke and Sailen Barik<br />

1. Introduction<br />

Since the advent of the polymerase chain reaction (PCR), a variety of PCR-based<br />

procedures of mutagenesis have been developed through the use of synthetic primers<br />

encoding the mutation. Among these, the megaprimer method and related ones (1–5)<br />

remain some of the simplest and most versatile. Variations and improvements of<br />

the basic technique have been suggested over the past few years; these include a<br />

combination of magapriming and overlap extension, improvement of yield, use of<br />

single-stranded DNA, avoidance of unwanted mutations arising from nontemplated<br />

insertions by Taq polymerase, and the inclusion of various kinds of mutations,<br />

including multiple, nonadjacent ones (2–11). The basic method (Fig. 1) requires three<br />

oligonucleotide primers and two PCRs (termed PCR-1 and -2 here) using the wild-type<br />

DNA as template (1,2,8,10). The “mutant” primer is represented by M and the two<br />

“outside” primers by A and B. The M primer may encode a substitution, a deletion, an<br />

insertion, or a combination of these mutations, thus providing versatility while using<br />

the same basic strategy (10). The first PCR (PCR-1) is performed using the mutant<br />

primer M and one of the outside primers, such as A (Fig. 1). The double-stranded<br />

product A-M is purified and used as a primer (hence the name megaprimer; ref. 1)<br />

in the second PCR (PCR-2) together with the other outside primer, B. Although both<br />

strands of the megaprimer may prime on the respective complementary strands of the<br />

template, the fundamental principles of PCR amplification ensure that only the one that<br />

extends to the other primer, that is, B in Fig. 1, will be exponentially amplified into<br />

the double stranded product in PCR-2. As mentioned, the wild-type DNA is used as<br />

template in both PCRs. This article describes the most optimized megaprimer method<br />

in our experience and has drawn freely on the improvements described by various<br />

authors (3–28).<br />

1.1. Improving the Yield of PCR-2<br />

Poor yields from PCR-2 have sometimes been reported even when proper primer<br />

design (see above) was followed, especially when the megaprimer is large (0.8 kb<br />

and above). Although the exact reasons remain unclear, the most likely reasons are<br />

From: Methods in Molecular <strong>Bio</strong>logy, Vol. 226: PCR Protocols, Second Edition<br />

Edited by: J. M. S. <strong>Bartlett</strong> and D. Stirling © Humana Press Inc., Totowa, NJ<br />

525


526 Burke and Barik<br />

Fig. 1. The basic megaprimer method. Primers A, B, M, and the priming strand of the<br />

megaprimer AM are indicated by thinner lines with arrowhead, while the thicker double lines<br />

represent the wild type template (usually part of a plasmid clone, not shown). Primers A and B<br />

contain restriction sites (e.g., NdeI and BamHI) indicated as thicker regions, and extra “clamp”<br />

sequence at the 5′ end indicated by double lines. The sequence to be inserted is shown as the<br />

dotted region in primer M and the subsequent PCR products. The final product containing the<br />

insertion is restricted and cloned.<br />

the unique features of the megaprimer, viz., its double-stranded nature and large size.<br />

Strand separation of the double-stranded megaprimer is essentially achieved in the<br />

denaturation steps of the PCR cycle. Under some conditions, however, self-annealing<br />

of the megaprimer apparently tends to reduce the yield of the product (4).<br />

Various solutions to this problem have been suggested. In one approach, a biotin tag<br />

is added to the 5′ end of primer A, which would generate a biotin-labeled megaprimer<br />

in PCR-1. After denaturation, the biotinylated strand of the megaprimer is purified on<br />

avidin attached to magnetic beads (26). In another method (27), the use of two<br />

parallel templates allowed the inclusion of two outside primers as well as the<br />

megaprimer in PCR-2, resulting in a direct amplification of the final product. Use of a<br />

“one-tube” method (described above), when properly optimized, should eliminate loss<br />

of megaprimer during the purification step. Other strategies for increasing the yield<br />

of PCR-2 involve optimizing the concentrations of both template and megaprimer. In


Mutagenesis by Megaprimer PCR 527<br />

some instances, the use of higher amounts of template (in the microgram range, as<br />

opposed to nanogram quantities used in standard PCR) in PCR-2 has been shown<br />

to dramatically increase the product yield (4). Unfortunately, higher concentrations<br />

of template also tend to increase mispriming by a megaprimer with a mismatched<br />

3′ end (our unpublished results). Thus, a more effective strategy may be to increase the<br />

amount of the megaprimer. A method that we have found useful is to perform the first<br />

several cycles of PCR-2 with the megaprimer only. After this initial asymmetric PCR,<br />

the small primer is added (11). In an optimization of this strategy (28), the starting<br />

concentration of megaprimer is increased to 6 µg (from 25 ng) per 100 µL of PCR-2.<br />

We have adopted a combination of the last two approaches in this article.<br />

2. Materials<br />

2.1. Template<br />

About 100 ng of DNA template to be mutated (e.g., a gene cloned in a plasmid).<br />

2.2. Primers<br />

100 pg of oligonucleotide primers A and B, and 50 ng of mutant primer M; one<br />

primer, say A, in the opposite sense, and other primer, B, in the same sense as the<br />

mutant primer M (Fig. 1). Include restriction sites, preferably unique, in these primers<br />

so that the final product can be efficiently digested with restriction enzymes and<br />

cloned. The mutant primer may be designed to contain a point mutation, or insertion,<br />

or deletion, as desired (see Notes 1 and 2).<br />

2.3. PCR Buffer<br />

10× PCR buffer for Pfu polymerase (Stratagene Cloning Systems, La Jolla, CA)<br />

is: 200 mM Tris-HCl (pH 8.0–8.3); 100 mM KCl; 20 mM MgCl 2 ; 60 mM ammonium<br />

sulfate; 1% Triton X-100 100 µg/mL nuclease-free BSA; the buffer is usually supplied<br />

with the enzyme by most manufacturers.<br />

2.4. Deoxyribonucleotides<br />

The final dNTP concentration is generally 200 µM for each nucleotide. Make a stock<br />

dNTP mix containing 2 mM of each dNTP (dATP, dCTP, dGTP, dTTP); we make it by<br />

adding 50 µL of 10 mM stock solutions of each nucleotide, available commercially,<br />

into 50 µL H 2 O, to produce 250 µL of the mix.<br />

2.5. Analysis and Purification of DNA<br />

A system for purifying the PCR products, such as gel electrophoresis, followed by<br />

recovery of the appropriate DNA band in the excised agarose fragment (8).<br />

Wherever needed in this procedure, use deionized (e.g., Millipore) autoclaved<br />

water.<br />

3. Method<br />

3.1. PCR-1: Synthesis of the Megaprimer<br />

1. It is assumed that the reader is familiar with standard PCR protocols. Use the following<br />

recipe for the first PCR. Make the following 100-µL reaction mix in an appropriate<br />

microcentrifuge tube (0.5 or 1.7 mL, dictated by the heating block of your thermal<br />

cycler): H 2 O (75 µL); 10× PCR buffer (10 µL); 2 mM each of dNTP mix (10 UL; the final


528 Burke and Barik<br />

concentration of each nucleotide is 200 µM ); Primer A (50 pmol); Primer M (50 pmol)<br />

(Note 3); DNA template (10–100 ng); and 2.5 U Pfu polymerase (0.5 µL);or 2.5 U Taq<br />

plus 0.1 U Pfu polymerase) for a total of 100 µL.<br />

2. Vortex well to mix, then spin briefly in a microfuge. If the thermal cycler has a heated lid,<br />

then proceed to do PCR; otherwise, reopen the tube, overlay the reaction mixture with<br />

enough mineral oil to cover the reaction (~100 µL for a 0.5-mL microfuge tube), then<br />

close cap. The tube is now ready for thermal cycling.<br />

3. Perform PCR-1 using the following cycle profiles. Initial denaturation: 94°C, 3 min;<br />

30 to 35 main cycles: 94°C, 1 min (denaturation): T° (depending on the T m of the primers),<br />

2 min (annealing); 72°C, appropriate time, depending on product length (extension); and<br />

final extension 72°C, 1.5 × N min.<br />

After synthesis, the samples are maintained at 4°C (called “soak” file in older Perkin–<br />

Elmer programs) for a specified time. Some instruments lack an active cooling mechanism<br />

and keep samples at an ambient temperature of about 20°C by circulating tap water around<br />

the heat block, which appears to be adequate for overnight runs; others just shut off at<br />

the end of the final extension.<br />

4. After PCR, proceed directly to the next step if there is no oil overlay. Otherwise, first<br />

remove the oil as follows. (If oil is not removed completely, the sample will float up when<br />

loaded in horizontal agarose gels!). Add 200 µL of chloroform to each tube. The mineral<br />

oil and chloroform will mix to form a single phase and sink to the bottom of the tube.<br />

Spin for 30 s in a microfuge. Carefully collect ~80 µL of top aqueous layer and transfer<br />

to a fresh Eppendorf tube.<br />

5. Purify the megaprimer using any standard procedures such as gel purification (see<br />

Chapter 18) and use it in PCR-2 below.<br />

3.2. PCR-2: Synthesis of the Mutant Using the Megaprimer<br />

1. Reconstitute 100-µL PCR as follows: 10× PCR buffer (10 µL); 2 mM each of dNTP mix<br />

(10 µL; final concentration of each nucleotide is 200 µM ); All of the recovered megaprimer<br />

(A-M) from the previous step (20–50 µL); DNA template (0.2 µg); Make up volume to<br />

100 µL with H 2 O; and mix well.<br />

2. Start reaction essentially as described for PCR-1, except that a “hot-start” is preferred (see<br />

Note 4) and is performed as follows. When the reaction is in the annealing step of the first<br />

cycle, open the cap briefly, quickly add 0.5 µL of Pfu polymerase (2.5 U, or 2.5 U Taq plus<br />

0.1 U Pfu polymerase), and mix by pipetting. Close the cap and let PCR continue.<br />

3. After five cycles, when the reaction is again at an annealing step, promptly add 50 pmol<br />

of primer B, mix well, and let PCR continue another 30 cycles. (The small amounts of<br />

primer B and Pfu polymerase do not contribute significantly to the total reaction volume<br />

and, therefore, have been ignored in the volume calculations).<br />

4. Do another PCR in parallel, using primers A and B (but no megaprimer) and the same<br />

wild-type template; use an aliquot (5 µL) of this PCR as a size marker when analyzing<br />

PCR-2 by gel electrophoresis. This will also help in identifying the real product (in PCR-2)<br />

among the wrong ones that sometimes result from mispriming.<br />

5. Gel purify the final mutant PCR product essentially as described earlier for the purification<br />

of the megaprimer (see Notes 5 and 6).<br />

4. Notes<br />

1. Design of the mutant primer. Perhaps the most unique feature of the megaprimer method<br />

is that the product of one PCR becomes a primer in the next, which creates the following<br />

potential problem. Taq polymerase, as a result of its lack of proofreading activity, tends


Mutagenesis by Megaprimer PCR 529<br />

to extend the product DNA beyond the template by adding one or two non-templated<br />

residues, predominantly As (12). When the product is used as a primer in the next<br />

round of PCR (PCR-2), these nontemplated A residues may not match with the template<br />

and, therefore, will either abrogate amplification (13–15) or produce an undesired<br />

A-substitution. A variety of solutions to this problem have been recommended (5,8,10,16).<br />

The first is to design the mutant primer such that there is at least one T residue beyond<br />

the 5′ end of the primer sequence in the template. Thus, when the complementary strand<br />

incorporates a non-templated A at the 3′ end, it will still be complementary to the other<br />

strand. If the template sequence does not permit this, a second solution is to use a mixture<br />

of Taq and Pfu DNA polymerases in 20:1 ratio in PCR-2 (3) or to use Pfu exclusively. This<br />

is what we have recommended in this chapter. The 3′ exonuclease activity of Pfu should<br />

remove any mismatch at the 3′ end of the megaprimer; however, this proofreading ability<br />

also necessitates the addition of at least 10 perfectly matched bases on both the 5′ and 3′<br />

ends of the mutagenic primer (8,10,13,17). Finally, one can use enzymes, such as mung<br />

bean nuclease, that will remove nontemplated nucleotides from the megaprimer (16).<br />

In addition to these unique considerations, the general rules of primer design described<br />

below, should be followed.<br />

2. Length of the megaprimer. Try to avoid making megaprimers (A-M) that approach the size<br />

of the final, full-length product (gene) A-B (see Fig. 1). Briefly, if M is too close to B, it<br />

will make separation of AB and AM (unincorporated, left-over megaprimer) difficult after<br />

PCR-2. When the mutation is to be created near B, one should make an M primer of the<br />

opposite polarity, and synthesize BM megaprimer (rather than AM), and then do PCR-2<br />

with BM megaprimer and A primer. When the mutation is at or very near the 5′ or 3′ end of<br />

the gene (within 1–50 nucleotides), there is no need to use the megaprimer method; one can<br />

simply incorporate the mutation in either A or B primer and do a straightforward PCR using<br />

A and B primers! For borderline situations, such as when the mutation is, for example,<br />

120 nucleotides away from the 5′ end of the gene, incorporation of the mutation in primer<br />

A may make the primer too big to synthesize; or else, it will make the megaprimer AM<br />

too short to purify away from primer B. In such a case, simply back up primer A to a few<br />

hundred bases further upstream to make the AM megaprimer longer. In general, realize<br />

that primers A and B can be located virtually anywhere on either side of the mutant<br />

primer M, and therefore, try to utilize this flexibility as an advantage when designing<br />

these primers.<br />

3. Molar amount of megaprimer. Because the megaprimer is large, one needs to use a greater<br />

quantity of it to achieve the same number of moles as a smaller primer. Example: 50 pmol<br />

of a 20 nt-long single-stranded primer will equal 0.3 µg; however, 50 pmol of a 500 nt-long<br />

double-stranded megaprimer will equal 6 µg. A good yield and recovery of megaprimer<br />

is, therefore, important. If needed, do 2× 100 µL PCRs to generate the megaprimer. There<br />

is no need to remove the template DNA after PCR-1 because the same DNA will be used<br />

as template in PCR-2.<br />

4. “Hot start” PCR-2. The hot-start technique used in PCR-2 works just as well as the more<br />

expensive commercial methods. Hot start tends to reduce false and nonspecific priming<br />

in PCR in general (29) and is particularly useful in PCR-2 of the megaprimer method<br />

(our unpublished observation).<br />

5. Poor yield of mutant. If the final yield is poor, the surest strategy is to amplify a portion<br />

of the gel-purified mutant product in a third PCR (PCR-3) using primers A and B and<br />

hot start. This may also be necessary if PCR-2 produces nonspecific products in addition<br />

to the specific one. Before PCR-3 is conducted, however, it is very important to ensure<br />

that the mutant product of PCR-2 is well separated from the wild type template in the<br />

gel purification; otherwise, PCR-3 will amplify the wild-type DNA as well. The final gel-


530 Burke and Barik<br />

purified mutant DNA (from either PCR-2 or PCR-3) is ready for a variety of applications,<br />

such as sequencing (30–36) or cloning (33,34).<br />

6. Single-tube methods. Recently, various investigators have reported successful modifications<br />

of the megaprimer method in which the purification step is not required. One involves<br />

cleavage of the template, coupled with enzymatic removal of PCR-1 primers, to ensure<br />

amplification of the correct product in PCR-2 (24). A second possibility is to exploit the<br />

unusually high T m of the megaprimer by designing a short, low Tm flanking primer for<br />

PCR-1, and a long flanking primer for PCR-2. This enables the use of a higher Tm for<br />

PCR-2 such that it will only allow annealing of the appropriate flanking primer (25).<br />

A third method uses a limiting amount of the first flanking primer, such that when the<br />

second flanking primer is added, the principle product will be the mutant DNA (17).<br />

Since we have not tested any of these modifications, the interested reader is advised to<br />

consult the original papers.<br />

Acknowledgments<br />

Research in the author’s laboratory was supported in part by NIH Grant AI45803.<br />

S. B. is also a recipient of a Burroughs Wellcome New Initiatives in Malaria Research<br />

Award.<br />

References<br />

1. Sarkar, G. and Sommer, S. S. (1990) The “megaprimer” method of site-directed mutagenesis.<br />

<strong>Bio</strong>techniques 8, 404– 407.<br />

2. Landt, O., Grunart, H.-P., and Hahn, U. (1990) A general method for rapid mutagenesis<br />

using the polymerase chain reaction. Gene 96, 125–128.<br />

3. Perrin, S. and Gilliland, G. (1990) Site-specific mutagenesis using asymmetric polymerase<br />

chain reaction and a single mutant primer. Nucleic Acids Res. 18, 7433–7438.<br />

4. Barik, S. and Galinski, M. (1991) “Megaprimer” method of PCR: Increased template<br />

concentration improves yield. <strong>Bio</strong>techniques 10, 489– 490.<br />

5. Kuipers, O. P., Boot, H. J., and de Vos, W. M. (1991) Improved site-directed mutagenesis<br />

method using PCR. Nucleic Acids Res. 19, 4558.<br />

6. Merino, E., Osuna, J., Bolivat, F., and Soberon, X. (1992) A general PCR-based method<br />

for single or combinatorial oligonucleotide-directed mutagenesis on pUC/M13 vectors.<br />

<strong>Bio</strong>Techniques 12, 508–510.<br />

7. Aiyar, A. and Leis, J. (1993) Modification of the megaprimer method of PCR mutagenesis:<br />

Improved amplification of the final product. <strong>Bio</strong>techniques 14, 366–369.<br />

8. Barik, S. (1993) Site-directed mutagenesis by double polymerase chain reaction. Methods<br />

Mol. <strong>Bio</strong>l. 15, 277–286.<br />

9. Pont-Kingdon, G. (1994) Construction of chimeric molecules by a two-step recombinant<br />

PCR method. <strong>Bio</strong>Techniques 16, 1010–1011.<br />

10. Barik, S. (1995) Site-directed mutagenesis by PCR: Substitution, insertion, deletion, and<br />

gene fusion. Methods Neurosci. 26, 309–323.<br />

11. Datta, A. K. (1995) Efficient amplificaton using ‘megaprimer’ by asymmetric polymerase<br />

chain reaction. Nucleic Acids Res. 23, 4530– 4531.<br />

12. Clark, J. M. (1988) Novel non-templated nucleotide addition reactions catalyzed by<br />

procaryotic and eucaryotic DNA polymerases. Nucleic Acids Res. 16, 9677–9686.<br />

13. Barnes, W. M. (1994) PCR amplification of up to 35-kb DNA with high fidelity and high<br />

yield from lambda bacteriophage templates. Proc. Natl. Acad. Sci. USA 91, 2216–2220<br />

14. Sarkar, G., Cassady, J., Bottema, C. D. K., and Sommer, S. S. (1990) Characterization of<br />

polymerase chain reaction amplification of specific alleles. Anal. <strong>Bio</strong>chem. 186, 64–68.


Mutagenesis by Megaprimer PCR 531<br />

15. Kwok, S., Kellogg, D. E., McKinney, N., Spasic, D., Goda, L., Levenson, C., et al.<br />

(1990) Effects of primer-template mismatches on the polymerase chain reaction: Human<br />

immunodeficiency virus type I model studies. Nucleic Acids Res. 18, 999–1005.<br />

16. Chattopadhyay D. and Raha, T. (1997) PCR mutagenesis: treatment of the megaprimer with<br />

mung bean nuclease improves yield. <strong>Bio</strong>techniques 6, 1054–1056.<br />

17. Picard, V., Ersdal-Badju, E., Lu, A., and Bock, S. C. (1994) A rapid and efficient one-tube<br />

PCR-based mutagenesis technique using Pfu DNA polymerase. Nucleic Acids Res. 22,<br />

2587–2591.<br />

18. Rychlik, W. (1995) Selection of primers for polymerase chain reaction. Mol. <strong>Bio</strong>technol.<br />

3, 129–134.<br />

19. New England <strong>Bio</strong>labs Catalog (1995) pp. 208–209.<br />

20. Innis, M. A., Myambo, K. B., Gelfand, D. H., and Brow, M. A. D. (1988) DNA sequencing<br />

with Thermus aquaticus DNA polymerase and direct sequencing of polymerase chain<br />

reaction-amplified DNA. Proc. Natl. Acad. Sci. USA 85, 9436–9440.<br />

21. Suggs, S. V., Hirose, T., Miyake, E. H., Kawashima, M. J., <strong>John</strong>son, K. I., and Wallace, R. B.<br />

(1981) Use of synthetic oligodeoxyribonucleotides for the isolation of specific cloned DNA<br />

sequences, in ICN-UCLA Symposium on Developmental <strong>Bio</strong>logy Using Purified Genes,<br />

Vol. 23 (Brown, D. D., ed.),. Academic Press, New York, pp. 683–693.<br />

22. Tautz, D. and Renz, M. (1983) An optimized freeze-squeeze method for recovering long<br />

DNA from agarose gels. Anal. <strong>Bio</strong>chem. 132, 14–19.<br />

23. Zintz, C. B. and Beebe, D. C. (1991) Rapid re-amplification of PCR products purified in<br />

low melting point agarose gels. <strong>Bio</strong>techniques 11, 158–162.<br />

24. Seraphin, B. and Kandels-Lewis, S. 1996. An efficient PCR mutagenesis strategy without<br />

gel purification step that is amenable to automation. Nucleic Acids Res. 24, 3276–3277.<br />

25. Ke, S. H. and Madison, E. L. (1997) Rapid and efficient site-directed mutagenesis by<br />

single-tube ‘megaprimer’ PCR method. Nucleic Acids Res. 25, 3371–3372.<br />

26. Fieschi, J., Niccoli, P., Camilla, C., Chames, P., Chartier, M., and Baty, D. (1996)<br />

Polymerase chain reaction-based site-directed mutagenesis using magnetic beads. Anal.<br />

<strong>Bio</strong>chem. 234, 210–214.<br />

27. Upender, M., Raj, L., and Weir, M. (1995) Megaprimer method for in vitro mutagenesis<br />

using parallel templates. <strong>Bio</strong>techniques 18, 29–32.<br />

28. Smith, A. M. and Klugman, K. P. (1997) “Megaprimer” method of PCR-based mutagenesis:<br />

the concentration of megaprimer is a critical factor. <strong>Bio</strong>techniques 22, 438– 442.<br />

29. Mullis, K. B. (1991) The polymerase chain reaction in an anemic mode: how to avoid cold<br />

oligodeoxyribonuclear fusion. PCR Methods Appl. 1, 1– 4.<br />

30. Meltzer, S. J. (1993) Direct sequencing of polymerase chain reaction products. Methods<br />

Mol. <strong>Bio</strong>l. 15, 137–142.<br />

31. Sarkar, G. and Bolander, M. E. (1995) Semi exponential cycle sequencing. Nucleic Acids<br />

Res. 23, 1269–1270.<br />

32. Maniatis, T., Fritsch, E. F., and Sambrook, J. (1982) Molecular Cloning: A Laboratory<br />

Manual. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY, pp. 113–119.<br />

33. Barik, S. (1993) Expression and biochemical properties of a protein Ser/Thr phosphatase<br />

encoded by bacteriophage λ. Proc. Natl. Acad. Sci. USA 15, 10,633–10,637.<br />

34. Mazumder, B., Adhikary, G., and Barik, S. (1994) Bacterial expression of human respiratory<br />

syncytial viral phosphoprotein P and identification of Ser237 as the site of phosphorylation<br />

by cellular casein kinase II. Virology 205, 93–103.<br />

35. Krowczynska, A. M. and Henderson, M. B. (1992) Efficient purification of PCR products<br />

using ultrafiltration. <strong>Bio</strong>techniques 13, 286–289.<br />

36. Stoflet, E. S., Koeberl, D. D., Sarkar, G., and Sommer, S. S. (1988) Genomic amplification<br />

with transcript sequencing. Science 239, 491– 494.


532 Burke and Barik

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!