Next Article in Journal
Multi-Faceted Roles of DNAJB Protein in Cancer Metastasis and Clinical Implications
Next Article in Special Issue
Understanding Life at High Temperatures: Relationships of Molecular Channels in Enzymes of Methanogenic Archaea and Their Growth Temperatures
Previous Article in Journal
Antidepressive Effect of Natural Products and Their Derivatives Targeting BDNF-TrkB in Gut–Brain Axis
Previous Article in Special Issue
Alicyclobacillus mali FL18 as a Novel Source of Glycosyl Hydrolases: Characterization of a New Thermophilic β-Xylosidase Tolerant to Monosaccharides
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Thermophilic Nucleic Acid Polymerases and Their Application in Xenobiology

MOE International Joint Research Laboratory on Synthetic Biology and Medicines, School of Biology and Biological Engineering, South China University of Technology, Guangzhou 510006, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2022, 23(23), 14969; https://doi.org/10.3390/ijms232314969
Submission received: 27 October 2022 / Revised: 22 November 2022 / Accepted: 27 November 2022 / Published: 29 November 2022
(This article belongs to the Special Issue Thermophilic and Hyperthermophilic Microbes and Enzymes 2.0)

Abstract

:
Thermophilic nucleic acid polymerases, isolated from organisms that thrive in extremely hot environments, possess great DNA/RNA synthesis activities under high temperatures. These enzymes play indispensable roles in central life activities involved in DNA replication and repair, as well as RNA transcription, and have already been widely used in bioengineering, biotechnology, and biomedicine. Xeno nucleic acids (XNAs), which are analogs of DNA/RNA with unnatural moieties, have been developed as new carriers of genetic information in the past decades, which contributed to the fast development of a field called xenobiology. The broad application of these XNA molecules in the production of novel drugs, materials, and catalysts greatly relies on the capability of enzymatic synthesis, reverse transcription, and amplification of them, which have been partially achieved with natural or artificially tailored thermophilic nucleic acid polymerases. In this review, we first systematically summarize representative thermophilic and hyperthermophilic polymerases that have been extensively studied and utilized, followed by the introduction of methods and approaches in the engineering of these polymerases for the efficient synthesis, reverse transcription, and amplification of XNAs. The application of XNAs facilitated by these polymerases and their mutants is then discussed. In the end, a perspective for the future direction of further development and application of unnatural nucleic acid polymerases is provided.

1. Introduction

Nucleic acid polymerases are enzymes that catalyze DNA or RNA synthesis, including DNA polymerases (DNAPs), RNA polymerases (RNAPs), reverse transcriptases (RTs), and RNA-dependent RNA polymerases (RdRps), which play central roles in the storage and transmission of genetic information in living organisms, and have been widely applied in molecular biology and biotechnology [1,2,3,4,5]. Their unique activities and functions have laid the foundation of many broadly used or modern techniques, including polymerase chain reaction (PCR), DNA sequencing, and DNA information storage [6,7,8,9]. Thermostability is a desired property of nucleic acid polymerases for many of their applications, especially those involving thermocycling. Therefore, nucleic acid polymerases derived from thermophilic microorganisms have been widely used in biotechnology, due to their innate tolerance towards high temperatures [10,11], and some of them have been further engineered to be hyperthermophilic for better performance in these applications [12].
Due to the limited number of building blocks, natural DNA and RNA have intrinsic constraints for their properties, functions, and applications, which in principle, can be expanded by incorporating more kinds of building blocks. The past few decades have witnessed the development of xeno nucleic acids (XNAs). The attempts to create these DNA and RNA analogs with unnatural moieties have fostered a new field named xenobiology. The term “xenobiology” derives from the Greek word xenos, which means “foreign, alien, or stranger” [13]. The main aim of xenobiology is to expand the framework of natural life forms with artificial building blocks, such as XNAs and non-canonical amino acids (ncAAs). These building blocks are orthogonal with natural components, which leads to increased genetic and functional diversity [14]. In addition, the new-to-nature feature of xenobiology prevents information exchange with natural systems, forming a “genetic firewall” [15].
Similar to DNA and RNA, XNAs also need to be enzymatically manipulated for broader application. For example, enzymatic synthesis and amplification of XNAs are essential for efficiently producing and evolving functional XNA molecules [16]. However, due to their exotic structures, the unnatural nucleoside triphosphates of many XNAs are poor substrates for natural nucleic acid polymerases, and thus employing protein engineering approaches to make efficient XNA polymerases (XNAPs) is one of the most urgent tasks in xenobiology, which has drawn broad research interest in recent years [17,18,19,20]. Many of the thermophilic nucleic acid polymerases have been used as the starting scaffolds for generating XNAPs, and their great thermostability is a useful feature for high-temperature synthesis and thermocycling amplification of XNAs [21]. With the engineered XNAPs, the great potential of XNAs in a broad range of applications has been extensively demonstrated [22,23].
In this review, we first summarize the thermophilic and hyperthermophilic nucleic acid polymerases that have been extensively studied and broadly applied in biotechnology. The strategies for engineering these polymerases to be efficient XNAPs are introduced next, followed by the summary of representative XNAPs, with the relationship of key mutations and activities discussed. The applications of XNAPs are then reviewed.

2. Thermophilic and Hyperthermophilic Nucleic Acid Polymerases

Generally speaking, extensively investigated and applied thermophilic and hyperthermophilic nucleic acid polymerases mainly include thermostable DNAPs and minority RNAPs mostly from Thermus, Thermococcus, and Pyrococcus [5,10,24] (Table 1). From the perspective of family classifications, the DNAPs among these enzymes are concentrated in family A and B (Figure 1).
Taq DNAP from the thermophilic bacterium Thermus aquaticus was the first isolated thermophilic DNAP [54], which led to a breakthrough in PCR technology by eliminating the addition of a new enzyme after each cycle of thermocycling [55,56]. The optimal temperature of Taq DNAP is 75 to 80 °C, which is much higher compared with DNAPs from organisms living in regular environments. However, its half-lives are 45 to 50 min at 95 °C and 9 min at 97.5 °C, which are relatively short [26]. Taq DNAP has been classified as family A. It has 5′-3′ exonuclease activity and no 3′-5′ exonuclease activity, so its fidelity is not good compared to polymerases that own 3′-5′ exonuclease activity [57]. Under optimized conditions, the error rate of Taq DNAP was tested to be about 1.2 × 10−5 to 3.3 × 10−6 mf × bp−1 × d−1 (mutation frequency per base pair per duplication) [25]. Based on the real-time PCR experiments, amplification efficiencies of Taq DNAP were found to be around 80% for targets shorter than 1 kb and around 60% for 2.6 kb targets with a CG content between 45 to 56% [58]. Efforts have been made to alter the properties of Taq DNAP to improve its performance for different applications. For example, the pH of the reaction buffer and MgCl2 concentration have been optimized to improve its fidelity [25]. Deletion of proper regions in the 5′-3′ exonuclease domain has proven effective in improving the fidelity or thermostability of Taq DNAP. KlenTaq, a truncated variant of Taq DNAP lacking the N-terminal 235 amino acids, has been reported to have a two-fold higher fidelity than that of Taq DNAP [59]. A similar variant, the Stoffel fragment (SF), which is deficient in the N-terminal 289 amino acids, was found to have an increased thermostability [26]. Besides the DNA amplification activity, Taq DNAP also demonstrated some extent of RNA RT activity [60]. With optimized conditions, Taq DNAP could facilitate one-enzyme reverse transcription-qPCR of viral RNA [61].
Similar A-family thermophilic DNAPs have been identified from other Thermus strains, such as Tfi DNAP from Thermus filiformis [62], Tth DNAP from Thermus thermophilus [28], Tfl DNAP from Thermus flavus [63], Tca DNAP from Thermus caldophilus [30], and TsK1 DNAP from Thermus scotoductus [31]. Like Taq DNAP, these DNAPs possess 5′-3′ exonuclease activity but no 3′-5′ exonuclease activity. The PCR performances, such as amplification efficiency, fidelity, and specificity, and reaction conditions of Tfi DNAP, are similar to those of Taq DNAP [27]. Removing the 5′-3′ exonuclease domain of Tfi DNAP did not significantly affect the enzyme activity and stability [64]. However, a comparative study of exo/exo+ Tfi DNAP blends with different blending ratios exhibited that raising the proportion of exo Tfi mutant led to an increase in the PCR amplification yield for the target product [65]. Tth DNAP from Thermus thermophilus HB8 also showed structural and functional similarities with Taq DNAP [66]. In addition, under similar conditions in the presence of Mn2+, Tth DNAP performed higher reverse transcription activity than Taq DNAP, which is useful for one-pot reverse transcription and PCR amplification of low-level RNA [67]. Sso7d is a small protein isolated from hyperthermophilic archaebacteria Saccharolobus solfataricus and may play a role of stabilizing the genomic DNA in the cell [68,69]. It has great thermostability and is able to bind with dsDNAs without much sequence preference [70]. An early study found that the fusion of this protein with several DNAPs significantly enhanced their processivity [71]. Recently, Sso7d protein was also fused to the N-terminal of Tth DNAP, which might improve the DNA binding capacity and processivity of Tth DNAP without affecting its catalytic activity and stability [66].
Though with a high sequence homology with Taq DNAP, some family A DNAPs from other Thermus strains exhibit some different characteristics. For example, Tfl DNAP from Thermus flavus demonstrated a higher thermostability and maintained PCR activity after heat treatment at 94 °C for 60 min, while Taq DNAP lost activity within 30 min under the same temperature [29]. Furthermore, Tfl and Tth DNAPs can significantly eliminate negative influences from the inhibitors of PCR reaction in the intraocular fluids and blood, avoiding false-negative results [72,73]. For another example, Tca DNAP exhibited longer half-lives in the presence of gelatin and a narrower working pH range than that of Taq DNAP [30,74]. Recently, a novel A-family DNAP, TsK1 DNAP, was reported to have a comparable half-life to rTaq (a commercially available recombinant Taq DNAP), which is shorter than that of Taq DNAP [31]. However, this enzyme demonstrated high amplification efficiency and better fidelity than Taq DNAP, making it a potential tool for molecular biology methodologies.
Many other family A DNAPs have been isolated from Bacillus species [75], such as Bst DNAP from Bacillus stearothermophilus [76] (now categorized as Geobacillus stearothermophilus [77]), Bca DNAP from Bacillus caldotenax [33], Bcav DNAP from Bacillus caldovelox [34], Bsm DNAP from Bacillus smithii [78], and Gss DNAP from Geobacillus sp. 777 [36]. The optimal temperatures of these DNAPs are 60 to 70 °C, which are lower than those of the thermostable polymerases from Thermus species introduced above. Bst-like DNAPs are widely used in isothermal amplification techniques, such as loop-mediated isothermal amplification (LAMP) and whole genome amplification (WGA), due to their strong strand displacement activity [79,80].
Hyperthermophilic microorganisms are bacteria or archaea whose optimal temperature for growth exceeds 80 °C [81]. Thermotoga, Thermosipho, Aquifex, and Thermocrinis are common genera of hyperthermophilic bacteria [82]. Tma DNAP isolated from Thermotoga maritima is a 97 kDa A-family polymerase with inherent 3′-5′ proofreading activity and 5′-3′ exonuclease activity [37]. Tma DNAP exhibited activity over a wide range of temperatures from 45 to 90 °C, with the optimal temperature being 75 to 80 °C. N-terminal truncation of Tma DNAP yielded UlTma (Perkin-Elmer) with enhanced thermostability [83]. The presence of 3′-5′ proofreading activity does not confer a high level of fidelity to UlTma, as implied by a similar replication accuracy with that of Taq DNAP [84]. A similar polymerase, Tne DNAP, has been isolated from Thermotoga neapolitana [38]. Later research found that mutations in the O-helix region improved the fidelity of this polymerase [39]. A mixture of KlenTaq and Tne DNAPs has also been prepared and found useful for the efficient amplification of long DNA fragments [85]. Aae DNAP isolated from Aquifex aeolicus is another family A DNAP, possessing 5′-3′ polymerase activity and 3′-5′ proofreading activity but no 5′-3′ exonuclease activity [40]. Half-lives of Aae DNAP, in the presence of BSA, were 6 h and 1.7 h at 75 and 85 °C, respectively. Although Aquifex aeolicus can grow at nearly 95 °C, the activity of Aae DNAP decreased rapidly at temperatures over 90 °C.
Family B DNAPs from hyperthermophilic archaea have been widely used in PCR due to their good thermostability and 3′-5′ proofreading activity [42,86]. Several thermostable DNAPs have been isolated from the genera Thermococcus and Pyrococcus, characterized, and commercialized. Tli (Vent) DNAP from Thermococcus litoralis is an archaeal DNAP, with a molecular weight of 89 kDa [41]. It is also the first reported thermostable DNAP possessing proofreading activity, which demonstrated a 2–4 times lower error rate compared to the proofreading activity-free enzyme Replinase DNAP (isolated from Thermus flavis) [41]. Tli DNAP is extremely thermostable, having a half-life of 2 h at 100 °C, and can be used for high-temperature DNA synthesis. In addition, Tli DNAP is resistant to hemoglobin inhibition, making it suitable for PCR amplification of DNAs in blood samples [72]. KOD DNAP is another commercial high-fidelity B-family polymerase possessing a 3′-5′ exonuclease domain and was isolated from Thermococcus kodakaraensis [42] (formerly Pyrococcus sp. KOD1 [87]). KOD DNAP has a higher thermostability than most DNAPs, and its half-life at 95 °C reaches 12 h. It has also been reported to have an extension rate of 6.0–7.8 kb/min and an error rate of 2.6 × 10−6, allowing efficient and faithful amplification of DNA in PCR reaction. PCR technique based on KOD DNAP was further developed for accurate amplification of long DNAs [88]. With a mixture of wild-type KOD DNAP and its exo variant (N210D), i.e., KOD Dash polymerase, long DNA fragments (up to 15 kb) were accurately amplified. 9°N DNAP, isolated from Thermococcus sp. 9°N-7, has a similar temperature-sensitive strand displacement activity and Km values with Tli DNAP [43]. Tgo DNAP, isolated from Thermococcus gorganarius, is another widely engineered polymerase for XNA synthesis [44]. Besides these DNAPs introduced above, many other B-family polymerases have also been isolated from Thermococcus species and characterized, including Tfu DNAP from Thermococcus fumicolans [45], TNA1 DNAP from Thermococcus sp. NA1 [46], Tpe DNAP from Thermococcus peptonophilus [47], Tzi DNAP from Thermococcus zilligii [48], and Twa DNAP from Thermococcus waiotapuensis [49].
Pfu DNAP is one of the most representative family B DNAPs isolated from hyperthermophilic marine archaea Pyrococcus furiosus [86]. Pfu DNAP has a high fidelity under optimized buffer and substrate concentrations [50]. The error rate of Pfu DNAP was found to be 1.38 × 10−6. This high fidelity is mainly attributed to the 3′-5′ exonuclease activity, and an exo mutant of Pfu DNAP demonstrated significantly decreased fidelity [50]. The extension rate of Pfu DNAP is only 0.5–1.5 kb/min, which is lower than that of most other DNAPs [10]. The fusion of the Sso7d protein to the N-terminal of Pfu DNAP improved its processivity but did not affect its catalytic activity and stability [71]. It was also found that Pfu DNAP has weak incorporation activity of dUTP, which reduced the PCR efficiency when dUTP was used in PCR reaction [89].
Pst DNAP, also known as Deep Vent DNAP, is another well-studied thermophilic archaeal DNAP. It was isolated from Pyrococcus strain GB-D, which can grow at 104 °C [90]. Pst DNAP possesses a high fidelity with an error rate of 2.7 × 10−6, better than that of Vent or Taq DNAP [50]. Like other B-family polymerases, Pst DNAP has a high 3′-5′ proofreading activity that decreases errors during the DNA replication process. Deletion of the 3′-5′ exonuclease activity also significantly reduced its fidelity [51].
Several other thermophilic DNAPs that are not as famous as Pfu and Deep Vent DNAPs have been isolated from other Pyrococcus species. For example, Pab DNAP, which is also a B-family DNAP, was isolated from Pyrococcus abyssi, an archaeon growing in hyperthermal environments in the deep sea [91]. Pab DNAP has a higher thermostability than Taq and Pfu DNAPs and retains 75% of its activity after being incubated at 100 °C for 5 h [52]. Pab DNAP also has 3′-5′ exonuclease activity that confers proofreading ability and high fidelity to it [52]. Another example is Pwo DNAP from Pyrococcus woesei [92]. This DNAP has a molecular weight of 90 kDa, and also possesses 3′-5′ exonuclease proofreading function like other B-family DNAPs [53]. For the highest activity, Pwo DNAP needs a slightly more alkaline buffer condition, which may lead to the degradation of dNTPs, and thus dNTPs should be added right before the addition of Pwo DNAP when preparing the PCR solution [93]. Besides, the 3′-5′ exonuclease activity of this polymerase can lead to the degradation of primers and PCR products when the concentrations of dNTPs are low, and nuclease-resistant phosphorothionate protected primers can be used to solve this problem [10].
All the thermophilic DNAPs introduced above are from family A and B, as summarized in Table 1. The phylogenetic tree of these polymerases shown in Figure 1 exhibits their evolutionary relationships. Although some of these DNAPs are neither well known nor commercialized, their identifications have expanded the repertoire of thermophilic DNAPs, providing more candidates to be explored and engineered for various potential applications.
Although natural thermophilic DNAPs are very efficient for DNA synthesis, and thus, have been broadly used in biotechnology, their activities of XNA synthesis are usually relatively poor, which severely limits their applications in xenobiology. In order to get efficient polymerases for the synthesis, reverse transcription, and even amplification or inter-transcription of XNAs (Figure 2), natural polymerases have to be engineered with various protein engineering strategies.

3. Strategies for Engineering Thermophilic Nucleic Acid Polymerases

The engineering of polymerases can be carried out via directed evolution, rational design, or semi-rational design [5,17,19,94]. Directed evolution mimics Darwinian evolution in nature, and yet with significantly shortened evolution time for desired phenotypic traits [95]. Random mutagenesis and/or recombination are carried out on the target polymerase genes with a much higher frequency than that of spontaneous mutagenesis or recombination in nature, followed by the selection or screening of desired mutants under artificial pressures. For polymerases with more structural information, rational or semi-rational approaches can be used to predict candidate residues or regions for mutagenesis, reducing the size of the polymerase library and the labor intensity for subsequent selection or screening [96].

3.1. Strategies for Mutant Generation or Library Construction

Error-prone PCR and DNA shuffling are two methods that are most extensively employed to randomize the gene of a target protein (Figure 3). Error-prone PCR is derived from standard PCR reaction, and yet polymerases of low fidelity and altered reaction conditions, including unbalanced concentrations of dNTPs and the addition of manganese ion are applied to increase the mutation rate of the target gene during amplification [97]. DNA shuffling provides a method to recombine homologous gene sequences, which is similar to natural homologous recombination but much more efficient [98]. Many strategies have been developed for DNA shuffling, including DNase I fragmentation and reassembly, staggered extension process (StEP), and synthetic shuffling. Traditional DNA shuffling involves DNase I digestion of a pool of homologous genes and subsequent reassembly of fragments by PCR [99]. Instead of random fragmentation and assembly, StEP uses the target genes as templates to create a recombinant library through multiple rounds of shortened polymerase-catalyzed extension [100]. In some cases, the addition of specific synthetic oligonucleotides during DNA shuffling can make the libraries more directional for studying the function of interest [101]. DNA shuffling usually requires high-sequence homology. For parental genes with insufficient homology, it may be a feasible method to optimize the shuffling template sequences through computer programs to improve the homology [102,103].
Rapid developments in sequencing, structure determination, and computational tools pave the way to the rational design of proteins. Through collection and analysis of existing sequence/structure-function data, candidate mutations of a protein for desired properties can be predicted. Site-directed mutagenesis is then carried out to generate target mutants or focused libraries. With a deeper understanding of sequence–structure-function relationships, even de novo design of proteins can be accomplished [104,105]. In past decades, many algorithms have been developed to facilitate structure prediction, design, and engineering of proteins, such as FoldX [106], Rosetta [107], I-Mutant [108], FRESCO [109], PROSS [110], and UniRep [111]. As an example of applying rational protein engineering approaches on polymerases, four Bst DNAP variants with enhanced thermostability have recently been obtained through MutCompute, an unsupervised machine learning algorithm [112].
Although having dramatically diminished the time and labor involved in the selection or screening of protein mutants, rational design requires extensive and in-depth data of sequence/structure-function relationships to improve the accuracy, which is not available for many proteins [113]. Combining the advantages of both directed evolution and rational design, semi-rational design has proven to be an effective tool for protein engineering. A small number of promising residues are identified based on computational simulation and analysis, leading to the construction of smaller but high-quality libraries and more efficient evolution processes. Various semi-rational approaches for protein engineering have been developed, including structure-based combinatorial protein engineering (SCOPE) [114], combinatorial active-site saturation test (CAST) [115], iterative saturation mutation (ISM) [116], sequence saturation mutagenesis (SeSaM) [117], protein sequence activity relationship algorithm (ProSAR) [118], and reconstructing evolutionary adaptive paths (REAP) [119]. Some of these approaches have been successfully used for the engineering of many thermophilic DNAPs, such as Bst DNAP and Taq DNAP [112,120].

3.2. Strategies for the Selection or Screening of Polymerase Libraries

Directed evolution is a powerful tool in the development of polymerases, in which the critical step is to build a high-throughput selection or screening method for the enrichment of active mutants. Selection or screening strategies for protein mutants are usually designed based on the binding of the proteins and their ligands, visualization of the catalytic activities of the enzymes, selective amplification of the target genes, or viability of the organisms [121]. The core for the selection or screening of a library is to connect the phenotypes of the mutants with their genotypes. At present, broadly used methods for selecting or screening polymerase mutants with unnatural activities mainly include methods based on the phage system, in vitro compartmentalization system, and multi-well plate system [17].
Based on the phage display technique, Romesberg and co-workers developed a polymerase selection system [120]. In this system, the polymerase library was co-displayed with the primer/template substrate on M13 phage particles, and successful extension of the primer with unnatural nucleoside triphosphates led to biotin labeling of the 3′-end of the primer, allowing the separation of active polymerase mutants from the library with streptavidin-coated beads. In another example, Liu and co-workers designed a phage-assisted continuous evolution (PACE) system, which could be used for iterative rounds of protein evolution without human intervention continuously [122]. This system correlated the desired activity of the target protein with the infectivity of the M13 phage and, thus, realized the rapid evolution of the protein along with the phage propagation.
In vitro compartmentalization is another strategy to build a linkage between genotypes and their corresponding phenotypes and has been broadly used in protein evolution. Some two decades ago, Tawfik and Griffiths developed “man-made cell-like compartments” using water-in-oil emulsions to generate separated micro-reactors, allowing the isolation of independent reactions and selection of promising protein variants [123]. Since then, emulsion-based compartmentalization has also been extensively applied to build polymerase evolution systems. Holliger and co-workers designed the compartmentalized self-replication (CSR) method based on microemulsion, in which polymerase variants were individually packaged into compartments of water-in-oil emulsion, together with PCR primers and nucleoside triphosphate substrates [124]. In this way, the genes of active variants were replicated by the polymerases that they encoded during thermocycling and enriched in the gene pool. Later, various derivative methods of CSR have been developed, including short-patch compartmentalized self-replication (spCSR) [125], reverse transcription-compartmentalized self-replication (RT-CSR) [126,127], compartmentalized partnered replication (CPR) [128], and high-temperature isothermal compartmentalized self-replication (HTI-CSR) [103].
To directly select or screen for hard-to-evolve polymerase mutants with XNA synthesis or reverse transcription activities, several other in vitro compartmentalization-based selection or screening strategies that do not rely on self-replication of the polymerase gene have been developed. For example, the compartmentalized self-tagging (CST) method was designed to select polymerases capable of XNA synthesis [129]. Similar to CSR, the polymerase pool was also compartmentalized with primers and nucleoside triphosphate substrates in water-in-oil emulsions to ensure the separation of individual variants and genotype–phenotype association. Different from CSR, CST was based on the tagging of a polymerase-encoding plasmid by extension of a short biotinylated primer when the polymerase had desired activity. Subsequent bead separation of the tagged plasmid allowed the enrichment of active polymerase variants. Later, to select for XNA RTs, Holliger and co-workers developed compartmentalized bead labeling (CBL), which relied on bead co-immobilization of the polymerase-encoding plasmids and primer/template complex for reverse transcription and subsequent fluorescent screening of the beads harboring desired variants [130].
To achieve a more controllable in vitro compartmentalization, microfluidic systems can be used to generate predefined compartments [131]. Recently, the Chaput group developed droplet-based optical polymerase sorting (DrOPS) method relying on microfluidic technology and cell sorting, in which polymerase variants were encapsulated with optical sensors for monitoring polymerase activity [132]. Successful extension of the primer by polymerase mutants led to the generation of fluorescence, and then, the water-in-oil-in-water or water-in-oil droplets were sorted by fluorescence-activated cell sorting (FACS) or fluorescence-activated droplet sorting (FADS) [133,134].
Although multi-well plate screening methods do not have throughputs as high as those of methods introduced above, they are still broadly used in the identification of polymerase variants from focused libraries with smaller size or from libraries pre-enriched with the methods introduced above [120,125,129,130]. In a typical multi-well plate screening method for polymerase mutants, the polymerase-mediated primer extension is correlated with the generation of colored or fluorescent products from enzymatic reactions, which can be directly monitored with a plate reader.

4. Thermophilic XNAPs

Thermophilic nucleic acid polymerases and their mutants have been extensively explored and used in the synthesis, reverse transcription, and even amplification of XNAs [5,135]. Although some other polymerases that are not high temperature tolerant, such as mutants of T7 RNAP, have also been used for the synthesis of modified nucleic acids [136], thermophilic nucleic acid polymerases are indispensable for XNA synthesis, reverse transcription, or amplification at high temperatures or with thermocycling programs, which are essential when tough templates with complex secondary structures are used or are important for higher yields.
Some thermophilic DNAPs, such as Taq, KlenTaq, Tth, KOD (exo), KOD Dash, Vent (exo), Pwo, Pfu, and Tgo DNAPs, demonstrate good tolerance to modifications on nucleobases [137,138,139,140,141,142] but are less tolerant to sugar modifications. However, syntheses or reverse transcriptions of different sugar-modified XNAs with limited lengths by certain polymerases have been reported, although not very efficient. For example, Taq DNAP was reported to be capable of reverse transcription or replication of hexose nucleic acid (HNA) with the length of a few nucleotides [143]. Bst DNAP has proven capable of reverse transcribing 2′-fluoro-arabino nucleic acid (FANA), α-L-threofuranosyl nucleic acid (TNA), and glycerol nucleic acid (GNA) [144,145]. Transcription or replication of short stretches of cyclohexenyl nucleic acid (CeNA) was demonstrated with Vent (exo) DNAP, and reverse transcription of short CeNA was realized with Taq DNAP or Vent (exo) DNAP [146]. Deep Vent (exo) is able to reverse transcribe short stretches of TNA templates into DNA, and effectively incorporate all four 2′-deoxy-2′-fluoro-β-D-arabinonucleoside 5′-triphosphates (2′-F-araNTPs) on a DNA template to yield full-length FANA products [147,148]. In general, most natural polymerases show relatively narrow substrate specificities and limited activities towards XNAs. This is likely due to the fact that, in nature, to play their respective roles, polymerases have to possess stringent substrate specificities to accurately discriminate the sugars (deoxyribose and ribose) in their substrates, so that they can use the correct template (DNA or RNA) and the correct nucleoside triphosphates (dNTPs or NTPs) for the synthesis of their target products [149]. The introduction of unnatural sugars into the nucleic acids also usually leads to a great change in their structures, which may contribute to their difficult recognition by the natural polymerases as well [150]. To overcome the stringent substrate specificity and increase the activity towards unnatural substrates, natural polymerases have to be engineered.
Some commercial polymerase mutants demonstrate enhanced synthesis efficiency for modified nucleic acids. For example, Therminator DNAP, a variant of 9°N (exo) DNAP, can incorporate various modified nucleotides [151,152,153] and even efficiently and faithfully synthesize long TNA oligonucleotides from DNA templates [153]. However, to realize efficient synthesis or reverse transcription of most of the fully substituted XNAs, the polymerases have to be further engineered via the directed evolution, rational design, or semi-rational design approaches summarized above.
Among thermophilic family A DNAPs, Taq DNAP and its truncated mutants, including SF and KlenTaq, are the most explored and engineered ones for expanded substrate repertoires. With a phage-display-based polymerase selection system, Romesberg and co-workers evolved an SF mutant, SFM19, that can incorporate 2′-O-methyl ribonucleoside triphosphates (2′-OMe-NTPs) on a DNA template [154]. Later, they further optimized the selection method and used SFM19 as the evolutionary starting point to evolve a series of polymerases, including SFM4-3, SFM4-6, and SFM4-9, that could transcribe or reverse transcribe fully 2′-OMe-modified oligonucleotides, or even PCR amplify partially 2′-OMe- or 2′-F-modified DNAs [120]. Further investigation demonstrated that these mutants could also synthesize or amplify other sugar-modified nucleic acids, including 2′-chloro (2′-Cl), 2′-amino (2′-Am), 2′-azido (2′-Az), and arabino-modified DNAs, and 2′-OMe- and 2′-F-modified RNAs [21,155,156]. Holliger and co-workers developed spCSR to select for variants of Taq DNAP, and obtained a mutant, AA40, with the ability to incorporate NTPs and sugar-modified nucleoside triphosphates [125].
Several thermophilic family B DNAPs have also been extensively engineered for the efficient synthesis and reverse transcription of various XNAs. For example, Holliger and co-workers used the CST method to select the libraries of TgoT DNAP (a Tgo DNAP mutant containing mutations V93Q, D141A, E143A, and A485L), and a mutant with HNA polymerase activity, Pol6G12, was obtained [129]. They also combined statistical correlation analysis (SCA) with activity screening or CST to develop a series of polymerases capable of synthesizing or reverse transcribing other XNAs, among which PolC7 can efficiently synthesize CeNA and LNA, PolD4K can efficiently synthesize ANA and FANA, RT521 can efficiently synthesize TNA and reverse-transcribe TNA, ANA, and FANA into DNA, while RT521K has good reverse transcription activity for CeNA and LNA. Later, by randomizing the positively charged and bulky residues of mutant RT521, which might lead to a steric clash between the polymerase surface and the P-ethyl-modification on the phosphate backbone, screening the libraries, and performing further site-directed mutagenesis, they successfully obtained mutant PGV2, which demonstrated substantially improved synthesis activity for a newly developed XNA with an uncharged backbone, alkyl phosphonate nucleic acids (phNA) [157]. Using CBL selection and plate-based screening methods, they further evolved a series of XNA RTs from mutant RT521K [130]. Among them, RT-TKK can efficiently reverse-transcribe D-altritol nucleic acid (AtNA). RT-C8, which was then evolved from RT-TKK, can efficiently reverse-transcribe 2′-OMe-RNA, and also has some extent of reverse transcription activity for P-α-S-phosphorothioate 2′-methoxyethyl RNA (PS 2′-MOE-RNA). Another derivative of RT-TKK, RT-H4, can reverse transcribe HNA much more efficiently than RT521K and RT-TKK.
Chaput and co-workers identified specificity-determining residues (SDRs) of the polymerase by analyzing the polymerase/DNA complex structure and screened for the beneficial mutations at SDR positions in a model polymerase scaffold [158]. By transferring these mutations to homologous proteins, a series of mutants that demonstrated RNA and TNA synthesis activities were rapidly developed from several family B DNAPs, including 9°N, Tgo, KOD, and Deep Vent DNAPs. They also successfully selected a manganese-independent TNA polymerase, 9n-YRI, from a site-saturation mutagenesis library of 9°N DNAP with the DrOPS method that they developed [132]. They further combined FADS sorting with deep mutational scanning to provide an unbiased screening of all possible single-point mutations in the finger subdomain of KOD (exo) DNAP [134]. By screening mutants containing combinations of selected mutations, a double mutant, KOD-RS, which can conduct efficient TNA synthesis, was obtained, suggesting that polymerase specificity may be controlled by a small number of highly specific residues and more attention should be paid to these sites when engineering polymerases for the synthesis of specific nucleic acids. They later developed a programmed allelic mutation (PAM) strategy, applied it with DrOPS sorting, and successfully evolved a mutant with enhanced efficiency and specificity for TNA synthesis, Kod-RSGA, from mutant Kod-RS [159]. Herdewijn and co-workers reported 3′-2′ phosphonomethylthreosyl nucleic acid (tPhoNA or PMT) as a novel genetic material, and carried out stepwise engineering of TgoT DNAP to produce a PMT polymerase [160]. By introducing mutations that are related to XNA synthesis activity and screening for mutations at key residues based on previously reported mutants, they successfully obtained mutant TgoT-EPFLH, which can efficiently synthesize PMT. They also demonstrated that PMT could be efficiently reverse transcribed into DNA by both TgoT mutant RT521 and KOD mutant K.RT521K. Based on structural analysis, Hoshino et al. developed variants of KOD DNAP for LNA synthesis and reverse transcription, among which KOD-DGLNK can efficiently synthesize LNA from DNA, and KOD-DLK can efficiently reverse transcribe LNA into DNA [161]. These two mutants also demonstrated transcription or reverse transcription activity for 2′-OMe-RNA, respectively. Recently, Chaput and co-workers systematically compared some of the representative XNAPs obtained in previous works introduced above and demonstrated their diversity in thermostability and activity, specificity, and fidelity for the synthesis or reverse transcription of different nucleic acids, including RNA, FANA, ANA, HNA, TNA, and PMT [162].

5. Key Mutations in Engineered XNAPs

The mutations of representative engineered thermophilic XNAPs introduced above are summarized in Table 2 and Figure 4, and the distribution of the mutation sites in the structures of these engineered XNAPs is illustrated in Figure 5. Among these mutations, some are crucial for the gain or enhancement of the activities towards unnatural substrates, based on the analysis of polymerase structures and testing of the effects of specific mutations.
Steric gate residues are crucial for DNAPs to discriminate the deoxyribose of dNTPs from the ribose of NTPs, and usually need to be engineered to allow the polymerases to adopt sugar-modified substrates well [168]. As an example for thermophilic family A DNAPs, residue E615 is the steric gate residue of Taq DNAP, and the side chain of E615 packs on the 2′ position of the sugar ring directly, which consequently hinders the incorporation of nucleotides with a larger group at the 2′ position [125]. Therefore, E615 and its adjacent residue I614 are frequently mutated in Taq DNAP or SF mutants that can recognize sugar-modified nucleoside triphosphates [154]. In SF mutant SFM19 (I614E, E615G), which can incorporate 2′-OMe-modified nucleotides, E615 was mutated to Gly, which has a much smaller side chain and facilitates the access of 2′-OMe-NTPs. The same mutation was harbored by Taq DNAP mutant AA40 (E602V, A608V, I614M, E615G), which exhibited the activity of incorporating nucleotides with a few 2′-modifications, such as 2′-F, 2′-N3 and 2′-OMe [125]. SFM19 derivatives SFM4-3 (SFM19: V518A, N583S, D655N, E681K, E742Q, M747R), SFM4-6 (SFM19: D655N, L657M, E681K, E742N, M747R) and SFM4-9 (SFM19: N415Y, V518A, D655N, L657M, E681V, E742N, M747R) acquired more mutations beyond the steric gate residue, which further increased their processivity and allowed them to efficiently transcribe, reverse transcribe, or even amplify 2′-modified nucleic acids [16,120,156]. Holliger and co-workers obtained two mutants of Taq DNAPs, M1 (G84A, D144G, K314R, E520G, F598L, A608V, E742G) and M4 (D58G, R74P, A109T, L245R, R343G, G370D, E520G, N583S, E694K, A743P), both of which contained the mutation of E520 to a smaller residue Gly, and acquired the ability for processing a diverse range of non-canonical substrates [163]. In principle, to bestow the polymerases with good reverse transcription or replication activities for XNAs, it is helpful to engineer their interaction with the template strand. Residues E742 and M747 are involved in the interaction of Taq DNAP with the template strand, and their mutations have been found to contribute to the enhanced reverse transcription activity of several Taq DNAP mutants by the introduction of a salt bridge between the side chain of the new amino acid and the phosphodiester group in the template or an increase in the positive charge [169,170,171]. Therefore, mutations at these two residues in SF mutants are likely important for the efficient reverse transcription and amplification of 2′-modified nucleic acids.
For the extensively explored and engineered thermophilic family B DNAPs, including 9°N, Tgo, and KOD DNAPs, mutations at several key residues are frequently included in mutants that are efficient for the synthesis of various XNAs [158]. Mutations D141A and E143A lead to the inactivation of the exonuclease activity of these polymerases, preventing the removal of the incorporated XNA nucleotides and, thus, are usually introduced into these enzymes before further engineering for better XNA synthesis activity. Besides D141A and E143A, mutation N210D of KOD DNAP can also make it deficient in exonuclease activity [88]. Mutation V93Q significantly reduces the uracil stalling of these polymerases and has been included in many mutants for XNA synthesis and reverse transcription [172]. Residue A485 was mutated to Leu in Therminator DNAP, which is a mutant of 9°N DNAP [153]. Mutation of A485 may affect the pocket shape of the active site and the substrate recognition [173] and is included in many efficient mutants for XNA synthesis. Similar to family A DNAPs, steric gate residues also play important roles in substrate discrimination in these family B DNAPs. Y409 and E664 have been identified as the steric-gate residue and the second steric-gate residue, respectively, of Tgo DNAP, and mutations of these two residues are usually found to be crucial for the efficient synthesis of RNA and many kinds of XNAs [164]. Residue I521 is close to the catalytic residue D542 and the active site of Tgo DNAP, and has been mutated to different amino acids in many mutants that demonstrate XNA RT activity, as well as several mutants that have good synthesis activity for some XNAs [129].
For 9°N DNAP, mutations at residues A485, Y409, and E664 significantly alter its TNA synthesis activity. For example, Therminator DNAP, which harbors mutation A485L, possesses high activity for TNA synthesis [174], and mutants 9n-YRI (D141A, E143A, A485R, E664I) and 9n-NVA (D141A, E143A, Y409N, D432G, A485V, V636A, E664A) are both efficient Mn2+-independent TNA polymerase with greatly enhanced fidelity [132].
Starting from Tgo DNAP mutant TgoT, which harbors mutations V93Q, D141A, E143A, and A485L, various XNAPs have been produced. Mutant TGK (TgoT: Y409G, E664K) has proven capable of synthesizing RNA, as well as pseudouridine-, 5-methyl-C-, 2′-F-, and 2′-Az-modified RNAs [164]. Besides these activities, mutant TGK also has the ability to synthesize short stretches of FANA, ANA, HNA, and TNA [162]. Mutant TGLLK (TgoT: Y409G, I521L, F545L, E664K) has the ability to incorporate 3′-deoxy- or 3′-O-methyl-NTPs to produce nucleic acids with 2′-5′ linkages [165]. Very recently, mutant 2M (TGLLK: T541G, K592A) and 3M (TGLLK: T541G, K592A, K664R), which included more mutations at the nascent-strand steric gate residues in the TGLLK scaffold, were found to be able to efficiently synthesize long MOE-RNA and 2′-OMe-RNA up to 750 nt [166]. Mutant Pol6G12 (TgoT: V589A, E609K, I610M, K659Q, E664Q, Q665P, R668K, D669Q, K671H, K674R, T676R, A681S, L704P, E730G) has proven an efficient HNA polymerase [129], and can also synthesize FANA [162]. Mutant PolC7 (TgoT: K659Q, V661A, E664Q, Q665P, D669A, K671Q, T676K, R709K) can synthesize CeNA and LNA. Mutant PolD4K (TgoT: L403P, P657T, E658Q, K659H, Y663H, E664K, D669A, K671N, T676I) is capable of synthesizing ANA and FANA, and is also able to synthesize TNA, HNA, PMT, and RNA [129,162]. Obviously, mutation at secondary steric gate residue E664 is critical for the synthesis activities of different XNAs, and appears in all these mutants of Tgo DNAP. A series of TgoT mutants harboring mutations at key residue I521 and many other residues have been selected for efficient reverse transcription of different XNAs [129]. Mutant RT521 (TgoT: E429G, I521L, K726R) is able to reverse-transcribe HNA, ANA, FANA, and TNA. Mutant RT521K contains additional mutations F445L and E664K than RT521, and is efficient for the reverse transcription of LNA and CeNA [130]. A range of RTs for tougher XNAs is derived from RT521K by including even more mutations. For example, mutant RT-TKK (RT521K: I114T, S383K, N735K) can efficiently reverse transcribe 2′-OMe-RNA and AtNA. Both residues S383 and N735 are mutated to a positively charged lysine, which possibly enhances the electrostatic interaction between the polymerase and the non-cognate template. Mutation I114T is located in the uracil-binding pocket and also contributes to the enhanced reverse transcription activity of 2′-OMe-RNA. Mutant RT-C8 (RT-TKK: F493V, Y496N, Y497L, Y499A, A500Q, K501H), which harbors many more mutations than RT-TKK, is even more efficient for the reverse transcription of 2′-OMe-RNA, and can also reverse transcribe 2′-MOE RNA and PS 2′-MOE-RNA. Another derivative mutant of RT-TKK, RT-H4 (RT-TKK: F493V, Y496H, Y497M, Y499F, A500E, K501N), can reverse transcribe HNA much more efficiently than RT-TKK. Reverse mutagenesis of the exonuclease activity-eliminating mutations of RT-C8 and RT-H4 led to the production of mutants RT-C8exo+ (RT-C8: A141D, A143E) and RT-H4exo+ (RT-H4: A141D, A143E), which restore the 3′-5′ exonuclease activity, and are the first XNA RTs with proofreading activity. Introduction of mutations P410T and S411R, which are located in the nucleotide-binding pocket based on the structure of highly homologous KOD DNAP, into RT521K generates mutant RT-TR (RT521K: P410T, S411R), which presents significantly improved fidelity of reverse transcription.
As introduced above, mutant 9n-YRI contains mutations A485R and E664I, which promote TNA synthesis in the absence of Mn2+. Due to the 91% sequence similarity between KOD and 9°N DNAPs, identical mutations have been transferred to KOD DNAP to produce a novel TNA polymerase. The resulting mutant Kod-RI (D141A, E143A, A485R, E664I) was also found to be an efficient TNA polymerase [158]. Furthermore, mutant Kod-RS (D141A, E143A, A485R, N491S) and Kod-QS (D141A, E143A, L489Q, N491S) showed reduced ability for DNA synthesis and stronger specificity for TNA substrates [134]. Chaput and coworkers speculated that the coordination between mutations A485R and N491S allows the polymerase to adapt to the structural changes of the non-cognate TNA/DNA duplex and the incoming TNA substrate. Mutant Kod-RSGA (D141A, E143A, A485R, N491S, R606G, T723A) exhibited even higher specificity for TNA substrates compared with Kod-RS, which revealed that R606G and T723A double mutations are contributing to converting KOD DNAP into a TNA polymerase [159]. To evaluate the effects of different mutations at key residues on the LNA synthesis performance of KOD DNAP, Hoshino et al. compared the LNA synthesis activity and fidelity of KOD mutants DVL, DVLK, DGLK, DVLNK, and DGLNK [161]. Other than Therminator mutation A485L and steric gate mutations Y409V (or Y409G) and E664K, D614N, which is located in the thumb subdomain and was found to contribute to the efficient synthesis of LNA and improved fidelity of incorporating 2′-OMe-NTPs, was also included in mutants KOD DVLNK and KOD DGLNK. Mutant KOD DGLNK (N210D, Y409G, A485L, D614N, E664K) was shown to be able to synthesize one kb long LNA and also efficiently synthesize 2′-OMe-RNA. In addition, both KOD DGLNK and KOD DLK (N210D, A485L, E664K) exhibited LNA reverse transcription activity, indicating that only three substitutions in KOD DNAP are required to make this polymerase an LNA reverse transcriptase. Notably, residue Y412 of Vent DNAP is homologous to the steric gate residue, Y409, of KOD DNAP, and mutation Y412G also reduces the substrate specificity of Vent DNAP and promotes the incorporation of 2′-modified nucleotides [175].

6. Application of XNAs and Thermophilic XNAPs

The unnatural moieties endow XNAs with expanded chemical and biological properties, which significantly broaden their application in various fields, spanning biotechnology, biomedicine, and nanotechnology [18] (Figure 6). The employment of efficient thermophilic XNAPs, either discovered or engineered, further facilitates the use of XNAs by allowing the transcription, reverse transcription, amplification, or evolution of them when needed [176].
Antisense oligonucleotides (ASOs) are short nucleic acids that can complementarily bind with target RNAs, such as mRNAs and miRNAs, to influence the expression of target genes through different mechanisms [177]. Development and exploration of chemically modified ASOs are urgently needed due to the limited properties of natural ASOs in nuclease resistance, target binding affinity, and pharmacokinetics. The first and the most broadly used chemical modification introduced into ASOs is phosphorothioate (PS), in which the non-bridging oxygen atom in the phosphodiester backbone is replaced by a sulfur atom [177,178]. The first ASO drug for treating cytomegalovirus retinitis approved by the US FDA, Vitravene (Fomivirsen), is PS modified [177,179]. However, the PS modification reduces its binding affinity to the target, resulting in an increased effective dose and accompanying toxicity [177,180,181,182]. Another ASO drug that has also received FDA approval is Kynamro (mipomersen sodium) [183]. It was modified with both PS and 2′-MOE to improve nuclease resistance and target RNA binding affinity [184]. In addition to 2′-MOE, FANA modification was also applied together with PS modification to produce ASOs with improved nuclease resistance and high binding affinity [185,186]. One of the most employed oligonucleotide modifications in recent years is LNA, which has a methylene bridge between the 2′-O and the 4′-C of the sugar ring [187]. It is characterized by a structure similar to RNA, high target binding affinity and specificity, high stability, and low toxicity, and thus excellent for ASO modification [188].
Ribozymes are catalytic RNA molecules that play important roles in many biochemical processes of living organisms, ranging from RNA splicing to protein synthesis [189]. Ribozymes and DNAzymes, which are catalytic DNA molecules, have found broad applications in biotechnology, biomedicine, biosensing, and biomaterials [190,191]. With the rise of XNAs, the possibility of creating XNAzymes as novel catalysts gradually attracted more and more attention [22]. Employing the XNAPs that they evolved from TgoT DNAP, Holliger and co-workers developed ANA, FANA, HNA, and CeNA XNAzymes with RNA endonuclease, RNA ligase, or XNA ligase activities [192]. Later, Chaput and co-workers evolved a FANAzyme that had a much higher catalytic rate for RNA cleavage, with different versions of Bst DNAP that can efficiently transcribe and reverse transcribe FANA [193]. By modifying classic DNAzyme 10-23 with FANA and TNA nucleotides, they also developed XNAzyme X10-23, which demonstrated good RNA cleavage activity and enhanced nuclease resistance while eliminating product inhibition [194]. In X10-23, the nucleotides in the substrate binding arms were fully substituted with FANA nucleotides, which in principle, would lead to higher substrate binding affinity due to the higher stability of FANA/RNA heteroduplexes compared with DNA/RNA heteroduplexes [195]. In addition, residues G2 and T8 in the catalytic core of X10-23 were also substituted with FANA nucleotides since it was found that these substitutions synergistically led to a 50% increase in the activity compared with the parental enzyme. Meanwhile, both the 5′ and 3′ ends of X10-23 were modified with a TNA nucleotide to protect it from nuclease degradation. With X10-23, they successfully silenced green fluorescent protein (GFP) and endogenous Kirsten rat sarcoma viral oncogene (KRAS) persistently. Recently, Yu and co-workers reported a TNA enzyme T8-6 that can catalyze the formation of 2′-5′ phosphate bonds in RNA ligation reactions, which was selected using Kod-RI DNAP [196]. In addition, they also developed a TNA enzyme Tz1, capable of cleaving RNA substrates, and successfully achieved target gene silencing in vivo [197]. The selection in vitro for this TNA enzyme was performed with Kod-RI DNAP and Bst 2.0 DNAP.
Clustered, regularly interspaced short palindromic repeats (CRISPR)/Cas9 system is a powerful tool for gene editing that has been extensively applied in biotechnology [198]. Recently, the introduction of unnatural modifications into guide RNAs (gRNAs) was found to be useful for improving the performance of CRISPR/Cas9 system. For example, Cleveland and co-workers reported a chemically modified CRISPR RNA (crRNA) with PS, 2′-F, 2′-OMe, and S-constrained ethyl (cEt) modifications, which demonstrated enhanced biostability and binding affinity to transactivating CRISPR RNA (tracrRNA), and improved activity in gene editing [199]. Ryan et al. suggested a method that greatly improves the specificity and reduces the off-target effect of the CRISPR/Cas9 system, in which 2′-O-methyl-3′-phosphonoacetate modifications were introduced into guide RNAs (gRNAs) at specific sites [200]. Sontheimer and co-workers introduced a set of crRNAs and tracrRNAs heavily or fully modified with 2′-F, 2′-OMe, or PS, which demonstrated good gene editing activity [201]. Another approach to improve the specificity of the CRISPR/Cas9 system via chemical modification of crRNA was proposed by Hubbard and co-workers [202]. They introduced bridged nucleic acids (2′,4′-BNANC [N-Me]) and LNA into specific sites of crRNAs to disrupt the binding of crRNAs to the off-target sequences, thus greatly improving the gene editing precision of CRISPR/Cas9 system.
Aptamers are single-stranded nucleic acid molecules that can bind specifically to target molecules and are usually selected from an oligonucleotide library through a process called systematic evolution of ligands by exponential enrichment (SELEX) [203]. Due to the limited biological stability and binding affinity of natural aptamers, many chemically modified aptamers have been developed. The first aptamer drug approved by FDA, Macugen (Pegaptanib sodium), which targets vascular endothelial growth factor (VEGF), is modified with 2′-F and 2′-OMe [204]. Using XNAPs to transcribe, reverse transcribe, or amplify XNA libraries, XNA aptamers can be directly evolved via SELEX. For example, Romesberg and co-workers evolved fully 2′-OMe-modified or partially 2′-F-modified aptamers targeting human neutrophil elastase (HNE) with SF mutants that they evolved [16,205]. Heemstra and co-workers evolved the first TNA aptamer targeting a small molecule, ochratoxin A (OTA), using KOD RI TNA polymerase and a large fragment of Bst DNAP I to transcribe and reverse transcribe the TNA library, and demonstrated its significant biostability and high specificity to the target [206]. Highly specific HIV-1 integrase-binding FANA aptamers with dissociation constants reaching picomolar levels have also been obtained via SELEX, in which the FANA library was created with mutant D4K of Tgo DNAP [207,208].
To use XNAs as materials for genetic information storage, propagation, and retrieval has always been a key focus of xenobiology, which is challenging, especially when implemented in vivo. As one example of the initial efforts, Matsuda and co-workers reported that DNA templates containing 4′-thio-dT and 4′-thio-dC modifications, which were prepared by PCR amplification with KOD dash DNAP, can be transcribed into functional RNAs in mammalian cells [209]. Pezo et al. demonstrated that DNA templates containing a few CeNA, AraNA, or HNA nucleotides could be recognized by DNAP in E. coli, achieving in vivo transfer of genetic information from these XNAs to DNA [210]. When a few nucleotides around the cysteine 146 codon of the thyA gene were replaced by CeNA, AraNA, or HNA, the thyA-gene-encoded thymidylate synthase was still functional in E. coli. In another work, PN-DNA, containing P3′-N5′ phosphoramidate bonds, was developed by Herdewijn and co-workers, and its building block, 5′-amino-2′,5′-deoxycytidine 5′-N-triphosphate (NH-dCTP), was successfully incorporated into the R67DHFR gene, which encodes the trimethoprim resistance, using Klenow fragment or Vent (exo) DNAP [211]. Through resistance screening, antibiotic-resistant colonies were obtained, implying PN-DNA is able to store genetic information in living cells. Another example of in vivo replication of chemically modified DNA comes from Tavassoli and colleagues, who reported successful expression of a green fluorescent protein gene, iLOV, that was constructed by click-linking oligonucleotides via 5′-azides and 3′-alkynes in E. coli [212].
In addition to the applications described above, XNAs are also widely used in biomaterials, DNA nanotechnology, biosensing, and other fields [18,156,213,214,215,216,217]. For example, Romesberg and co-workers used SFM4-3 polymerase to PCR amplify DNA containing 2′-Az-A for preparing novel hydrogels [154]. Holliger and co-workers employed their evolved XNAPs to synthesize four different XNA strands and assembled them into XNA nanostructures [218]. With the continuous development of novel XNAs and their polymerases, XNAs will find ever-increasing applications, resulting in further expansion of the scope and increase of the importance of xenobiology.

7. Conclusions and Perspective

Since their isolation from organisms thriving at high temperatures, thermophilic nucleic acid polymerases have found broad application in almost all areas of biology and biotechnology [11]. The use of these enzymes is the core of many fundamental enabling techniques, including PCR and DNA sequencing [1,219]. Most of these enzymes that have been extensively explored and applied belong to family A and B of DNAPs, and their properties and activities varied significantly with their sources, structures, and functions in their original hosts, which leads to their diverse application scopes [219].
The development of XNAs to expand the scope of genetic materials is one of the core aspects of xenobiology. XNAs are derived from DNA and RNA, but display greater diversity in properties and functions, due to the unnatural moieties introduced into their structural units [220,221]. Polymerases that can efficiently transcribe, reverse transcribe, or amplify XNAs are essential for the full play of XNA functions. Despite their broad application, most of the natural nucleic acid polymerases demonstrate relatively poor activity towards unnatural substrates, which impedes their direct use as XNAPs. While a few thermophilic DNAPs were found to have polymerase activities for limited kinds of XNAs, many more efficient polymerases for various XNAs have been produced by engineering several well-studied thermophilic DNAPs with the approaches of directed evolution, rational design, or semi-rational design [18,96,150]. The discovery of these XNAPs immediately led to many applications of XNAs, such as the production of stable aptamers and XNAzymes [16,222].
Although current XNAPs already possess unnatural activities that allow some applications of XNAs, more efforts are still needed to improve their properties and performances. For example, for many XNA polymerases, the XNA synthesis or reverse transcription efficiencies and fidelities are still not satisfactory, and the lengths of the XNA products are still limited. Polymerases that can directly replicate and even PCR amplify fully modified XNAs have to be developed for better and broader use of XNAs in the future. Undoubtedly, the rapid development of novel protein design and engineering strategies, exemplified by machine learning approaches, will significantly facilitate our efforts on further engineering existing XNAPs to be better ones and, thus, broaden the application of XNAs [113]. Furthermore, the efforts on developing XNAPs have so far been mainly focused on engineering family A and B DNAPs from thermophiles, and thermophilic polymerases of other families may also be explored and engineered for XNA synthesis, reverse transcription, and amplification activities in the future, which may provide us more and even better XNAPs for different application scenarios. Hopefully, XNAPs that perform as well as natural DNA and RNA polymerases can be produced in the future, which will make XNAs become genetic materials comparable to, and even much better in some aspects, than DNA and RNA, and contribute to bringing xenobiology into an unprecedentedly thriving era.

Author Contributions

Conceptualization, T.C., G.W. and Y.D.; writing—original draft preparation, G.W., X.M., F.Y., Y.Q., Y.W., Y.X. and R.T.; writing—review and editing, T.C., G.W. and Y.D.; supervision, T.C.; funding acquisition, T.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (21978100), the National Key R&D Program of China (2019YFA0904102), the Guangdong Provincial Pearl River Talents Program (2019QN01Y228), and the Program for Guangdong Introducing Innovative and Entrepreneurial Teams (2019ZT08Y318).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

The authors thank the National Natural Science Foundation of China (21978100), the National Key R&D Program of China (2019YFA0904102), the Guangdong Provincial Pearl River Talents Program (2019QN01Y228), and the Program for Guangdong Introducing Innovative and Entrepreneurial Teams (2019ZT08Y318) for the financial support of this work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Aschenbrenner, J.; Marx, A. DNA Polymerases and Biotechnological Applications. Curr. Opin. Biotechol. 2017, 48, 187–195. [Google Scholar] [CrossRef]
  2. Borkotoky, S.; Murali, A. The Highly Efficient T7 RNA Polymerase: A Wonder Macromolecule in Biological Realm. Int. J. Biol. Macromol. 2018, 118, 49–56. [Google Scholar] [CrossRef] [PubMed]
  3. Martin-Alonso, S.; Frutos-Beltran, E.; Menendez-Arias, L. Reverse Transcriptase: From Transcriptomics to Genome Editing. Trends Biotechnol. 2021, 39, 194–210. [Google Scholar] [CrossRef]
  4. Venkataraman, S.; Prasad, B.V.L.S.; Selvarajan, R. RNA Dependent RNA Polymerases: Insights from Structure, Function and Evolution. Viruses 2018, 10, 76. [Google Scholar] [CrossRef] [Green Version]
  5. Laos, R.; Thomson, J.M.; Benner, S.A. DNA Polymerases Engineered by Directed Evolution to Incorporate Non-Standard Nucleotides. Front. Microbiol. 2014, 5, 565. [Google Scholar] [CrossRef] [Green Version]
  6. Zhu, H.L.; Zhang, H.Q.; Xu, Y.; Lassakova, S.; Korabecna, M.; Neuzil, P. PCR Past, Present and Future. Biotechniques 2020, 69, 317–325. [Google Scholar] [CrossRef]
  7. Roberts, M.A.J. Recombinant DNA Technology and DNA Sequencing. Essays Biochem. 2019, 63, 457–468. [Google Scholar] [CrossRef]
  8. Shendure, J.; Balasubramanian, S.; Church, G.M.; Gilbert, W.; Rogers, J.; Schloss, J.A.; Waterston, R.H. DNA Sequencing at 40: Past, Present and Future. Nature 2017, 550, 345–353. [Google Scholar] [CrossRef] [PubMed]
  9. Meiser, L.C.; Nguyen, B.H.; Chen, Y.J.; Nivala, J.; Strauss, K.; Ceze, L.; Grass, R.N. Synthetic DNA Applications in Information Technology. Nat. Commun. 2022, 13, 352. [Google Scholar] [CrossRef]
  10. Terpe, K. Overview of Thermostable DNA Polymerases for Classical PCR Applications: From Molecular and Biochemical Fundamentals to Commercial Systems. Appl. Microbiol. Biotechnol. 2013, 97, 10243–10254. [Google Scholar] [CrossRef]
  11. Ishino, S.; Ishino, Y. DNA Polymerases as Useful Reagents for Biotechnology—The History of Developmental Research in the Field. Front. Microbiol. 2014, 5, 465. [Google Scholar] [CrossRef] [Green Version]
  12. Kim, K.P.; Cho, S.S.; Lee, K.K.; Youn, M.H.; Kwon, S.T. Improved Thermostability and PCR Efficiency of Thermococcus celericrescens DNA Polymerase via Site-Directed Mutagenesis. J. Biotechnol. 2011, 155, 156–163. [Google Scholar] [CrossRef] [PubMed]
  13. Budisa, N.; Kubyshkin, V.; Schmidt, M. Xenobiology: A Journey towards Parallel Life Forms. ChemBioChem 2020, 21, 2228–2231. [Google Scholar] [CrossRef]
  14. Kubyshkin, V.; Budisa, N. Synthetic Alienation of Microbial Organisms by Using Genetic Code Engineering: Why and How? Biotechnol. J. 2017, 12, 1600097. [Google Scholar] [CrossRef]
  15. Schmidt, M. Xenobiology: A New Form of Life as the Ultimate Biosafety Tool. Bioessays 2010, 32, 322–331. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Thirunavukarasu, D.; Chen, T.; Liu, Z.X.; Hongdilokkul, N.; Romesberg, F.E. Selection of 2′-Fluoro-Modified Aptamers with Optimized Properties. J. Am. Chem. Soc. 2017, 139, 2892–2895. [Google Scholar] [CrossRef] [PubMed]
  17. Chen, T.; Romesberg, F.E. Directed Polymerase Evolution. FEBS Lett. 2014, 588, 219–229. [Google Scholar] [CrossRef] [Green Version]
  18. Zhu, G.; Song, P.; Wu, J.; Luo, M.; Chen, Z.; Chen, T. Application of Nucleic Acid Frameworks in the Construction of Nanostructures and Cascade Biocatalysts: Recent Progress and Perspective. Front. Bioeng. Biotechol. 2022, 9, 792489. [Google Scholar] [CrossRef]
  19. Sun, L.; Ma, X.; Zhang, B.; Qin, Y.; Ma, J.; Du, Y.; Chen, T. From Polymerase Engineering to Semi-Synthetic Life: Artificial Expansion of the Central Dogma. RSC Chem. Biol. 2022, 3, 1173–1197. [Google Scholar] [CrossRef]
  20. Hervey, J.R.D.; Freund, N.; Houlihan, G.; Dhaliwal, G.; Holliger, P.; Taylor, A.I. Efficient Synthesis and Replication of Diverse Sequence Libraries Composed of Biostable Nucleic Acid Analogues. RSC Chem. Biol. 2022, 10, 1209–1215. [Google Scholar] [CrossRef]
  21. Chen, T.; Romesberg, F.E. Polymerase Chain Transcription: Exponential Synthesis of RNA and Modified RNA. J. Am. Chem. Soc. 2017, 139, 9949–9954. [Google Scholar] [CrossRef]
  22. Taylor, A.I.; Holliger, P. Directed Evolution of Artificial Enzymes (XNAzymes) from Diverse Repertoires of Synthetic Genetic Polymers. Nat. Protoc. 2015, 10, 1625–1642. [Google Scholar] [CrossRef] [PubMed]
  23. Yang, K.F.; McCloskey, C.M.; Chaput, J.C. Reading and Writing Digital Information in TNA. Acs Synth. Biol. 2020, 9, 2936–2942. [Google Scholar] [CrossRef] [PubMed]
  24. Date, T.; Suzuki, K.; Imahori, K. Purification and Some Properties of DNA-Dependent RNA Polymerase from an Extreme Thermophile, Thermus thermophilus HB8. J. Biochem. 1975, 78, 845–858. [Google Scholar] [CrossRef] [PubMed]
  25. Eckert, K.A.; Kunkel, T.A. High Fidelity DNA Synthesis by the Thermus aquaticus DNA Polymerase. Nucleic Acids Res. 1990, 18, 3739–3744. [Google Scholar] [CrossRef] [Green Version]
  26. Lawyer, F.C.; Stoffel, S.; Saiki, R.K.; Chang, S.Y.; Landre, P.A.; Abramson, R.D.; Gelfand, D.H. High-Level Expression, Purification, and Enzymatic Characterization of Full-Length Thermus aquaticus DNA Polymerase and a Truncated form Deficient in 5′ to 3′ Exonuclease Activity. PCR Methods Appl. 1993, 2, 275–287. [Google Scholar] [CrossRef] [Green Version]
  27. Choi, J.J.; Jung, S.E.; Kim, H.K.; Kwon, S.T. Purification and Properties of Thermus filiformis DNA Polymerase Expressed in Escherichia coli. Biotechnol. Appl. Biochem. 1999, 30, 19–25. [Google Scholar]
  28. Dabrowski, S.; Kur, J. Recombinant His-Tagged DNA Polymerase. I. Cloning, Purification and Partial Characterization of Thermus thermophilus Recombinant DNA Polymerase. Acta Biochim. Pol. 1998, 45, 653–660. [Google Scholar] [CrossRef] [Green Version]
  29. Harrell, R.A.; Hart, R.P. Rapid Preparation of Thermus flavus DNA Polymerase. PCR Methods Appl. 1994, 3, 372–375. [Google Scholar] [CrossRef]
  30. Park, J.H.; Kim, J.S.; Kwon, S.T.; Lee, D.S. Purification and Characterization of Thermus caldophilus GK24 DNA Polymerase. Eur. J. Biochem. 1993, 214, 135–140. [Google Scholar] [CrossRef]
  31. Saghatelyan, A.; Panosyan, H.; Trchounian, A.; Birkeland, N.K. Characteristics of DNA Polymerase I from an Extreme Thermophile, Thermus scotoductus Strain K1. MicrobiologyOpen 2021, 10, e1149. [Google Scholar] [CrossRef] [PubMed]
  32. Ignatov, K.B.; Barsova, E.V.; Fradkov, A.F.; Blagodatskikh, K.A.; Kramarova, T.V.; Kramarov, V.M. A Strong Strand Displacement Activity of Thermostable DNA Polymerase Markedly Improves the Results of DNA Amplification. Biotechniques 2014, 57, 81–87. [Google Scholar] [CrossRef] [PubMed]
  33. Uemori, T.; Ishino, Y.; Fujita, K.; Asada, K.; Kato, I. Cloning of the DNA Polymerase Gene of Bacillus caldotenax and Characterization of the Gene Product. J. Biochem. 1993, 113, 401–410. [Google Scholar] [CrossRef]
  34. Sellmann, E.; Schröder, K.L.; Knoblich, I.M.; Westermann, P. Purification and Characterization of DNA Polymerases from Bacillus Species. J. Bacteriol. 1992, 174, 4350–4355. [Google Scholar] [CrossRef] [Green Version]
  35. Andrianova, M.; Komarova, N.; Grudtsov, V.; Kuznetsov, E.; Kuznetsov, A. Amplified Detection of the Aptamer-Vanillin Complex with the Use of Bsm DNA Polymerase. Sensors 2018, 18, 49. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Oscorbin, I.P.; Boyarskikh, U.A.; Filipenko, M.L. Large Fragment of DNA Polymerase I from Geobacillus sp 777: Cloning and Comparison with DNA Polymerases I in Practical Applications. Mol. Biotechnol. 2015, 57, 947–959. [Google Scholar] [CrossRef]
  37. Gelfand, D.H.; Lawyer, F.C. DNA Encoding a Thermostable Nucleic Acid Polymerase Enzyme from Thermotoga maritima. U.S. Patent No. 5,374,553, 20 December 1994. [Google Scholar]
  38. Slater, M.R.; Huang, F.; Hartnett, J.R.; Bolchakova, E.; Storts, D.R.; Otto, P.; Miller, K.M.; Novikov, A.; Velikodvorskaya, G.A. Thermophilic DNA Polymerases from Thermotoga neapolitana. U.S. Patent No. 6,077,664, 20 June 2000. [Google Scholar]
  39. Yang, S.W.; Astatke, M.; Potter, J.; Chatterjee, D.K. Mutant Thermotoga neapolitana DNA Polymerase I: Altered Catalytic Properties for Non-Templated Nucleotide Addition and Incorporation of Correct Nucleotides. Nucleic Acids Res. 2002, 30, 4314–4320. [Google Scholar] [CrossRef] [Green Version]
  40. Chang, J.R.; Choi, J.J.; Kim, H.K.; Kwon, S.T. Purification and Properties of Aquifex aeolicus DNA Polymerase Expressed in Escherichia coli. FEMS Microbiol. Lett. 2001, 201, 73–77. [Google Scholar] [CrossRef]
  41. Mattila, P.; Korpela, J.; Tenkanen, T.; Pitkänen, K. Fidelity of DNA Synthesis by the Thermococcus litoralis DNA Polymerase—An Extremely Heat Stable Enzyme with Proofreading Activity. Nucleic Acids Res. 1991, 19, 4967–4973. [Google Scholar] [CrossRef]
  42. Takagi, M.; Nishioka, M.; Kakihara, H.; Kitabayashi, M.; Inoue, H.; Kawakami, B.; Oka, M.; Imanaka, T. Characterization of DNA Polymerase from Pyrococcus sp. Strain KOD1 and Its Application to PCR. Appl. Environ. Microbiol. 1997, 63, 4504–4510. [Google Scholar] [CrossRef] [Green Version]
  43. Southworth, M.W.; Kong, H.; Kucera, R.B.; Ware, J.; Jannasch, H.W.; Perler, F.B. Cloning of Thermostable DNA Polymerases from Hyperthermophilic Marine Archaea with Emphasis on Thermococcus sp. 9 Degrees N-7 and Mutations Affecting 3′-5′ Exonuclease Activity. Proc. Natl. Acad. Sci. USA 1996, 93, 5281–5285. [Google Scholar] [CrossRef]
  44. Hopfner, K.P.; Eichinger, A.; Engh, R.A.; Laue, F.; Ankenbauer, W.; Huber, R.; Angerer, B. Crystal Structure of a Thermostable Type B DNA Polymerase from Thermococcus gorgonarius. Proc. Natl. Acad. Sci. USA 1999, 96, 3600–3605. [Google Scholar] [CrossRef] [Green Version]
  45. Cambon-Bonavita, M.A.; Schmitt, P.; Zieger, M.; Flaman, J.M.; Lesongeur, F.; Raguénès, G.; Bindel, D.; Frisch, N.; Lakkis, Z.; Dupret, D.; et al. Cloning, Expression, and Characterization of DNA Polymerase I from the Hyperthermophilic Archaea Thermococcus fumicolans. Extremophiles 2000, 4, 215–225. [Google Scholar] [CrossRef]
  46. Kim, Y.J.; Lee, H.S.; Bae, S.S.; Jeon, J.H.; Lim, J.K.; Cho, Y.; Nam, K.H.; Kang, S.G.; Kim, S.J.; Kwon, S.T.; et al. Cloning, Purification, and Characterization of a New DNA Polymerase from a Hyperthermophilic Archaeon, Thermococcus sp. NA1. J. Microbiol. Biotechnol. 2007, 17, 1090–1097. [Google Scholar]
  47. Lee, J.I.; Kim, Y.J.; Bae, H.; Cho, S.S.; Lee, J.H.; Kwon, S.T. Biochemical Properties and PCR Performance of a Family B DNA Polymerase from Hyperthermophilic Euryarchaeon Thermococcus peptonophilus. Appl. Biochem. Biotechnol. 2010, 160, 1585–1599. [Google Scholar] [CrossRef]
  48. Griffiths, K.; Nayak, S.; Park, K.; Mandelman, D.; Modrell, B.; Lee, J.; Ng, B.; Gibbs, M.D.; Bergquist, P.L. New High Fidelity Polymerases from Thermococcus species. Protein Expr. Purif. 2007, 52, 19–30. [Google Scholar] [CrossRef]
  49. Cho, S.S.; Kim, K.P.; Lee, K.K.; Youn, M.H.; Kwon, S.T. Characterization and PCR Application of a New High-Fidelity DNA Polymerase from Thermococcus waiotapuensis. Enzym. Microb. Technol. 2012, 51, 334–341. [Google Scholar] [CrossRef]
  50. Cline, J.; Braman, J.C.; Hogrefe, H.H. PCR Fidelity of pfu DNA Polymerase and Other Thermostable DNA Polymerases. Nucleic Acids Res. 1996, 24, 3546–3551. [Google Scholar] [CrossRef] [Green Version]
  51. Huang, H.; Keohavong, P. Fidelity and Predominant Mutations Produced by Deep Vent Wild-Type and Exonuclease-Deficient DNA Polymerases during in Vitro DNA Amplification. DNA Cell Biol. 1996, 15, 589–594. [Google Scholar] [CrossRef]
  52. Dietrich, J.; Schmitt, P.; Zieger, M.; Preve, B.; Rolland, J.L.; Chaabihi, H.; Gueguen, Y. PCR Performance of the Highly Thermostable Proof-Reading B-Type DNA Polymerase from Pyrococcus abyssi. FEMS Microbiol. Lett. 2002, 217, 89–94. [Google Scholar] [CrossRef]
  53. Ghasemi, A.; Salmanian, A.H.; Sadeghifard, N.; Salarian, A.A.; Gholi, M.K. Cloning, Expression and Purification of Pwo Polymerase from Pyrococcus woesei. Iran. J. Microbiol. 2011, 3, 118–122. [Google Scholar] [PubMed]
  54. Chien, A.; Edgar, D.B.; Trela, J.M. Deoxyribonucleic Acid Polymerase from the Extreme Thermophile Thermus aquaticus. J. Bacteriol. 1976, 127, 1550–1557. [Google Scholar] [CrossRef] [PubMed]
  55. Mullis, K.; Faloona, F.; Scharf, S.; Saiki, R.; Horn, G.; Erlich, H. Specific Enzymatic Amplification of DNA In Vitro: The Polymerase Chain Reaction. Cold Spring Harb. Symp. Quant. Biol. 1986, 51, 263–273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Saiki, R.K.; Gelfand, D.H.; Stoffel, S.; Scharf, S.J.; Higuchi, R.; Horn, G.T.; Mullis, K.B.; Erlich, H.A. Primer-Directed Enzymatic Amplification of DNA with a Thermostable DNA Polymerase. Science 1988, 239, 487–491. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Tindall, K.R.; Kunkel, T.A. Fidelity of DNA Synthesis by the Thermus aquaticus DNA Polymerase. Biochemistry 1988, 27, 6008–6013. [Google Scholar] [CrossRef] [PubMed]
  58. Arezi, B.; Xing, W.; Sorge, J.A.; Hogrefe, H.H. Amplification Efficiency of Thermostable DNA Polymerases. Anal. Biochem. 2003, 321, 226–235. [Google Scholar] [CrossRef]
  59. Barnes, W.M. The Fidelity of Taq Polymerase Catalyzing PCR Is Improved by an N-Terminal Deletion. Gene 1992, 112, 29–35. [Google Scholar] [CrossRef]
  60. Grabko, V.I.; Chistyakova, L.G.; Lyapustin, V.N.; Korobko, V.G.; Miroshnikov, A.I. Reverse Transcription, Amplification and Sequencing of Poliovirus RNA by Taq DNA Polymerase. FEBS Lett. 1996, 387, 189–192. [Google Scholar] [CrossRef] [Green Version]
  61. Bhadra, S.; Maranhao, A.C.; Paik, I.; Ellington, A.D. One-Enzyme Reverse Transcription qPCR Using Taq DNA Polymerase. Biochemistry 2020, 59, 4638–4645. [Google Scholar] [CrossRef]
  62. Jung, S.E.; Choi, J.J.; Kim, H.K.; Kwon, S.T. Cloning and Analysis of the DNA Polymerase-Encoding Gene from Thermus filiformis. Mol. Cells. 1997, 7, 769–776. [Google Scholar]
  63. Akhmetzjanov, A.A.; Vakhitov, V.A. Molecular Cloning and Nucleotide Sequence of the DNA Polymerase Gene from Thermus flavus. Nucleic Acids Res. 1992, 20, 5839. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Choi, J.J.; Kim, H.K.; Kwon, S.T. Purification and Characterization of the 5′→3′ Exonuclease Domain-Deleted Thermus filiformis DNA Polymerase Expressed in Escherichia coli. Biotechnol. Lett. 2001, 23, 1647–1652. [Google Scholar] [CrossRef]
  65. Zheng, W.; Lee, J.E.; Potter, R.J.; Mandelman, D. DNA Polymerase Blends and Mutant DNA Polymerases. U.S. Patent Application No 11/170,762, 28 December 2006. [Google Scholar]
  66. Aye, S.L.; Fujiwara, K.; Ueki, A.; Doi, N. Engineering of DNA Polymerase I from Thermus thermophilus Using Compartmentalized Self-Replication. Biochem. Biophys. Res. Commun. 2018, 499, 170–176. [Google Scholar] [CrossRef] [PubMed]
  67. Myers, T.W.; Gelfand, D.H. Reverse Transcription and DNA Amplification by a Thermus thermophilus DNA Polymerase. Biochemistry 1991, 30, 7661–7666. [Google Scholar] [CrossRef] [PubMed]
  68. Choli, T.; Henning, P.; Wittmann-Liebold, B.; Reinhardt, R. Isolation, Characterization and Microsequence Analysis of a Small Basic Methylated DNA-Binding Protein from the Archaebacterium, Sulfolobus solfataricus. Biochim. Biophys. Acta. 1988, 950, 193–203. [Google Scholar] [CrossRef] [PubMed]
  69. Sakai, H.D.; Kurosawa, N. Saccharolobus caldissimus Gen. nov., sp nov., a Facultatively Anaerobic Iron-Reducing Hyperthermophilic Archaeon Isolated from an Acidic Terrestrial Hot Spring, and Reclassification of Sulfolobus solfataricus as Saccharolobus solfataricus comb. Nov and Sulfolobus shibatae as Saccharolobus shibatae comb. nov. Int. J. Syst. Evol. Micr. 2018, 68, 1271–1278. [Google Scholar]
  70. Gao, Y.G.; Su, S.Y.; Robinson, H.; Padmanabhan, S.; Lim, L.; McCrary, B.S.; Edmondson, S.P.; Shriver, J.W.; Wang, A.H. The Crystal Structure of the Hyperthermophile Chromosomal Protein Sso7d Bound to DNA. Nat. Struct. Biol. 1998, 5, 782–786. [Google Scholar] [CrossRef]
  71. Wang, Y.; Prosen, D.E.; Mei, L.; Sullivan, J.C.; Finney, M.; Vander Horn, P.B. A Novel Strategy to Engineer DNA Polymerases for Enhanced Processivity and Improved Performance in Vitro. Nucleic Acids Res. 2004, 32, 1197–1207. [Google Scholar] [CrossRef]
  72. Al-Soud, W.A.; Radstrom, P. Purification and Characterization of PCR-Inhibitory Components in Blood Cells. J. Clin. Microbiol. 2001, 39, 485–493. [Google Scholar] [CrossRef] [Green Version]
  73. Wiedbrauk, D.L.; Werner, J.C.; Drevon, A.M. Inhibition of PCR by Aqueous and Vitreous Fluids. J. Clin. Microbiol. 1995, 33, 2643–2646. [Google Scholar] [CrossRef] [Green Version]
  74. Kwon, S.T.; Kim, J.S.; Park, J.H.; Kim, H.K.; Lee, D.S. Cloning and Analysis of the DNA Polymerase-Encoding Gene from Thermus caldophilus GK24. Mol. Cells. 1997, 7, 264–271. [Google Scholar] [PubMed]
  75. Perler, F.B.; Kumar, S.; Kong, H. Thermostable DNA Polymerases. Adv. Protein Chem. 1996, 48, 377–435. [Google Scholar] [PubMed]
  76. Aliotta, J.M.; Pelletier, J.J.; Ware, J.L.; Moran, L.S.; Benner, J.S.; Kong, H. Thermostable Bst DNA Polymerase I Lacks a 3′→5′ Proofreading Exonuclease Activity. Genet. Anal. Biomol. Eng. 1996, 12, 185–195. [Google Scholar] [CrossRef]
  77. Nazina, T.N.; Tourova, T.P.; Poltaraus, A.B.; Novikova, E.V.; Grigoryan, A.A.; Ivanova, A.E.; Lysenko, A.M.; Petrunyaka, V.V.; Osipov, G.A.; Belyaev, S.S.; et al. Taxonomic Study of Aerobic Thermophilic Bacilli: Descriptions of Geobacillus subterraneus Gen. nov., sp. nov. and Geobacillus uzenensis sp. nov. from Petroleum Reservoirs and Transfer of Bacillus stearothermophilus, Bacillus thermocatenulatus, Bacillus thermoleovorans, Bacillus kaustophilus, Bacillus thermodenitrificans to Geobacillus as the New Combinations G. stearothermophilus, G. thermocatenulatus, G. thermoleovorans, G. kaustophilus, G. thermoglucosidasius and G. thermodenitrificans. Int. J. Syst. Evol. Microbiol. 2001, 51, 433–446. [Google Scholar]
  78. Hayashizaki, Y.; Itoh, M.; Benno, Y.; Lezhava, A. Novel DNA Polymerase. U.S. Patent Application No. 20100047862A1, 25 February 2010. [Google Scholar]
  79. Notomi, T.; Okayama, H.; Masubuchi, H.; Yonekawa, T.; Watanabe, K.; Amino, N.; Hase, T. Loop-Mediated Isothermal Amplification of DNA. Nucleic Acids Res. 2000, 28, e63. [Google Scholar] [CrossRef]
  80. Oscorbin, I.P.; Belousova, E.A.; Boyarskikh, U.A.; Zakabunin, A.I.; Khrapov, E.A.; Filipenko, M.L. Derivatives of Bst-Like Gss-Polymerase with Improved Processivity and Inhibitor Tolerance. Nucleic Acids Res. 2017, 45, 9595–9610. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Kelly, R.M.; Adams, M.W. Metabolism in Hyperthermophilic Microorganisms. Antonie Van Leeuwenhoek 1994, 66, 247–270. [Google Scholar]
  82. Stetter, K.O. A Brief History of the Discovery of Hyperthermophilic Life. Biochem. Soc. Trans. 2013, 41, 416–420. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Gelfand, D.H.; Lawyer, F.C.; Stoffel, S. Mutated Thermostable Nucleic Acid Polymerase Enzyme from Thermotoga maritima. U.S. Patent No. 5,420,029, 30 May 1995. [Google Scholar]
  84. Diaz, R.S.; Sabino, E.C. Accuracy of Replication in the Polymerase Chain Reaction. Comparison between Thermotoga maritima DNA Polymerase and Thermus aquaticus DNA Polymerase. Brazilian J. Med. Biol. Res. 1998, 31, 1239–1242. [Google Scholar] [CrossRef] [Green Version]
  85. Davalieva, K.G.; Efremov, G.D. A New Thermostable DNA Polymerase Mixture for Efficient Amplification of Long DNA Fragments. Appl. Biochem. Microbiol. 2010, 46, 230–234. [Google Scholar] [CrossRef]
  86. Lundberg, K.S.; Shoemaker, D.D.; Adams, M.W.; Short, J.M.; Sorge, J.A.; Mathur, E.J. High-Fidelity Amplification Using a Thermostable DNA Polymerase Isolated from Pyrococcus furiosus. Gene 1991, 108, 1–6. [Google Scholar] [CrossRef] [PubMed]
  87. Atomi, H.; Fukui, T.; Kanai, T.; Morikawa, M.; Imanaka, T. Description of Thermococcus kodakaraensis sp. nov., a Well Studied Hyperthermophilic Archaeon Previously Reported as Pyrococcus sp. KOD1. Archaea 2004, 1, 263–267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Nishioka, M.; Mizuguchi, H.; Fujiwara, S.; Komatsubara, S.; Kitabayashi, M.; Uemura, H.; Takagi, M.; Imanaka, T. Long and Accurate PCR with a Mixture of KOD DNA Polymerase and Its Exonuclease Deficient Mutant Enzyme. J. Biotechnol. 2001, 88, 141–149. [Google Scholar] [CrossRef]
  89. Slupphaug, G.; Alseth, I.; Eftedal, I.; Volden, G.; Krokan, H.E. Low Incorporation of dUMP by Some Thermostable DNA Polymerases May Limit Their Use in PCR Amplifications. Anal. Biochem. 1993, 211, 164–169. [Google Scholar] [CrossRef] [PubMed]
  90. Jannasch, H.W.; Wirsen, C.O.; Molyneaux, S.J.; Langworthy, T.A. Comparative Physiological Studies on Hyperthermophilic Archaea Isolated from Deep-Sea Hot Vents with Emphasis on Pyrococcus Strain GB-D. Appl. Environ. Microbiol. 1992, 58, 3472–3481. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Gueguen, Y.; Rolland, J.L.; Lecompte, O.; Azam, P.; Le Romancer, G.; Flament, D.; Raffin, J.P.; Dietrich, J. Characterization of Two DNA Polymerases from the Hyperthermophilic Euryarchaeon Pyrococcus abyssi. Eur. J. Biochem. 2001, 268, 5961–5969. [Google Scholar] [CrossRef]
  92. Dabrowski, S.; Kur, J. Cloning and Expression in Escherichia coli of the Recombinant His-Tagged DNA Polymerases from Pyrococcus furiosus and Pyrococcus woesei. Protein Expr. Purif. 1998, 14, 131–138. [Google Scholar] [CrossRef]
  93. Hidajat, R.; McNicol, P. Primer-Directed Mutagenesis of an Intact Plasmid by Using Pwo DNA Polymerase in Long Distance Inverse PCR. Biotechniques 1997, 22, 32–34. [Google Scholar] [CrossRef]
  94. Coulther, T.A.; Stern, H.R.; Beuning, P.J. Engineering Polymerases for New Functions. Trends Biotechnol. 2019, 37, 1091–1103. [Google Scholar] [CrossRef]
  95. Romero, P.A.; Arnold, F.H. Exploring Protein Fitness Landscapes by Directed Evolution. Nat. Rev. Mol. Cell Biol. 2009, 10, 866–876. [Google Scholar] [CrossRef] [Green Version]
  96. Nikoomanzar, A.; Chim, N.; Yik, E.J.; Chaput, J.C. Engineering Polymerases for Applications in Synthetic Biology. Q. Rev. Biophys. 2020, 53, e8. [Google Scholar] [CrossRef] [PubMed]
  97. McCullum, E.O.; Williams, B.A.; Zhang, J.L.; Chaput, J.C. Random Mutagenesis by Error-Prone PCR. Methods Mol. Biol. 2010, 634, 103–109. [Google Scholar] [PubMed]
  98. Stemmer, W.P. Rapid Evolution of a Protein in Vitro by DNA Shuffling. Nature 1994, 370, 389–391. [Google Scholar] [CrossRef] [PubMed]
  99. Stemmer, W.P. DNA Shuffling by Random Fragmentation and Reassembly: In Vitro Recombination for Molecular Evolution. Proc. Natl. Acad. Sci. USA 1994, 91, 10747–10751. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Zhao, H.; Giver, L.; Shao, Z.; Affholter, J.A.; Arnold, F.H. Molecular Evolution by Staggered Extension Process (StEP) In Vitro Recombination. Nat. Biotechnol. 1998, 16, 258–261. [Google Scholar] [CrossRef]
  101. Ness, J.E.; Kim, S.; Gottman, A.; Pak, R.; Krebber, A.; Borchert, T.V.; Govindarajan, S.; Mundorff, E.C.; Minshull, J. Synthetic Shuffling Expands Functional Protein Diversity by Allowing Amino Acids to Recombine Independently. Nat. Biotechnol. 2002, 20, 1251–1255. [Google Scholar] [CrossRef]
  102. Milligan, J.N.; Garry, D.J. Shuffle Optimizer: A Program to Optimize DNA Shuffling for Protein Engineering. Methods Mol. Biol. 2017, 1472, 35–45. [Google Scholar]
  103. Milligan, J.N.; Shroff, R.; Garry, D.J.; Ellington, A.D. Evolution of a Thermophilic Strand-Displacing Polymerase Using High-Temperature Isothermal Compartmentalized Self-Replication. Biochemistry 2018, 57, 4607–4619. [Google Scholar] [CrossRef] [Green Version]
  104. Huang, P.S.; Boyken, S.E.; Baker, D. The Coming of Age of De Novo Protein Design. Nature 2016, 537, 320–327. [Google Scholar] [CrossRef]
  105. Marcos, E.; Silva, D.A. Essentials of De Novo Protein Design: Methods and Applications. Wires Comput. Mol. Sci. 2018, 8, e1374. [Google Scholar] [CrossRef]
  106. Guerois, R.; Nielsen, J.E.; Serrano, L. Predicting Changes in the Stability of Proteins and Protein Complexes: A Study of More than 1000 Mutations. J. Mol. Biol. 2002, 320, 369–387. [Google Scholar] [CrossRef]
  107. Rohl, C.A.; Strauss, C.E.M.; Misura, K.M.S.; Baker, D. Protein Structure Prediction Using Rosetta. Methods Enzymol. 2004, 383, 66–93. [Google Scholar]
  108. Capriotti, E.; Fariselli, P.; Casadio, R. I-Mutant2.0: Predicting Stability Changes upon Mutation from the Protein Sequence or Structure. Nucleic Acids Res. 2005, 33, W306–W310. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Wijma, H.J.; Floor, R.J.; Jekel, P.A.; Baker, D.; Marrink, S.J.; Janssen, D.B. Computationally Designed Libraries for Rapid Enzyme Stabilization. Protein Engi. Des. Sel. 2014, 27, 49–58. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Goldenzweig, A.; Goldsmith, M.; Hill, S.E.; Gertman, O.; Laurino, P.; Ashani, Y.; Dym, O.; Unger, T.; Albeck, S.; Prilusky, J.; et al. Automated Structure- and Sequence-Based Design of Proteins for High Bacterial Expression and Stability. Mol. Cell 2016, 63, 337–346. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Alley, E.C.; Khimulya, G.; Biswas, S.; AlQuraishi, M.; Church, G.M. Unified Rational Protein Engineering with Sequence-Based Deep Representation Learning. Nat. Methods 2019, 16, 1315–1322. [Google Scholar] [CrossRef]
  112. Paik, I.; Ngo, P.H.T.; Shroff, R.; Diaz, D.J.; Maranhao, A.C.; Walker, D.J.F.; Bhadra, S.; Ellington, A.D. Improved Bst DNA Polymerase Variants Derived via a Machine Learning Approach. Biochemistry 2021. [Google Scholar] [CrossRef]
  113. Yang, K.K.; Wu, Z.; Arnold, F.H. Machine-Learning-Guided Directed Evolution for Protein Engineering. Nat. Methods 2019, 16, 687–694. [Google Scholar] [CrossRef]
  114. O’Maille, P.E.; Bakhtina, M.; Tsai, M.D. Structure-Based Combinatorial Protein Engineering (SCOPE). J Mol Biol. 2002, 321, 677–691. [Google Scholar] [CrossRef]
  115. Reetz, M.T.; Bocola, M.; Carballeira, J.D.; Zha, D.; Vogel, A. Expanding the Range of Substrate Acceptance of Enzymes: Combinatorial Active-Site Saturation Test. Angew. Chem. Int. Edit. 2005, 44, 4192–4196. [Google Scholar] [CrossRef]
  116. Reetz, M.T.; Carballeira, J.D. Iterative Saturation Mutagenesis (ISM) for Rapid Directed Evolution of Functional Enzymes. Nat. Protoc. 2007, 2, 891–903. [Google Scholar] [CrossRef] [PubMed]
  117. Wong, T.S.; Tee, K.L.; Hauer, B.; Schwaneberg, U. Sequence Saturation Mutagenesis (SeSaM): A Novel Method for Directed Evolution. Nucleic Acids Res. 2004, 32, e26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Fox, R.J.; Davis, S.C.; Mundorff, E.C.; Newman, L.M.; Gavrilovic, V.; Ma, S.K.; Chung, L.M.; Ching, C.; Tam, S.; Muley, S.; et al. Improving Catalytic Function by ProSAR-Driven Enzyme Evolution. Nat. Biotechnol. 2007, 25, 338–344. [Google Scholar] [CrossRef] [PubMed]
  119. Cole, M.F.; Gaucher, E.A. Exploiting Models of Molecular Evolution to Efficiently Direct Protein Engineering. J. Mol. Evol. 2011, 72, 193–203. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Chen, T.; Hongdilokkul, N.; Liu, Z.; Adhikary, R.; Tsuen, S.S.; Romesberg, F.E. Evolution of Thermophilic DNA Polymerases for the Recognition and Amplification of C2′-Modified DNA. Nat. Chem. 2016, 8, 556–562. [Google Scholar] [CrossRef]
  121. Markel, U.; Essani, K.D.; Besirlioglu, V.; Schiffels, J.; Streit, W.R.; Schwaneberg, U. Advances in Ultrahigh-Throughput Screening for Directed Enzyme Evolution. Chem. Soc. Rev. 2020, 49, 233–262. [Google Scholar]
  122. Esvelt, K.M.; Carlson, J.C.; Liu, D.R. A System for the Continuous Directed Evolution of Biomolecules. Nature 2011, 472, 499–503. [Google Scholar] [CrossRef] [Green Version]
  123. Tawfik, D.S.; Griffiths, A.D. Man-Made Cell-Like Compartments for Molecular Evolution. Nat. Biotechnol. 1998, 16, 652–656. [Google Scholar] [CrossRef]
  124. Ghadessy, F.J.; Ong, J.L.; Holliger, P. Directed Evolution of Polymerase Function by Compartmentalized Self-Replication. Proc. Natl. Acad. Sci. USA 2001, 98, 4552–4557. [Google Scholar] [CrossRef]
  125. Ong, J.L.; Loakes, D.; Jaroslawski, S.; Too, K.; Holliger, P. Directed Evolution of DNA Polymerase, RNA Polymerase and Reverse Transcriptase Activity in a Single Polypeptide. J. Mol. Biol. 2006, 361, 537–550. [Google Scholar] [CrossRef]
  126. Ellefson, J.W.; Gollihar, J.; Shroff, R.; Shivram, H.; Iyer, V.R.; Ellington, A.D. Synthetic Evolutionary Origin of a Proofreading Reverse Transcriptase. Science 2016, 352, 1590–1593. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Shroff, R.; Ellefson, J.W.; Wang, S.S.; Boulgakov, A.A.; Hughes, R.A.; Ellington, A.D. Recovery of Information Stored in Modified DNA with an Evolved Polymerase. ACS Synth. Biol. 2022, 11, 554–561. [Google Scholar] [CrossRef] [PubMed]
  128. Ellefson, J.W.; Meyer, A.J.; Hughes, R.A.; Cannon, J.R.; Brodbelt, J.S.; Ellington, A.D. Directed Evolution of Genetic Parts and Circuits by Compartmentalized Partnered Replication. Nat. Biotechnol. 2014, 32, 97–101. [Google Scholar] [CrossRef] [PubMed]
  129. Pinheiro, V.B.; Taylor, A.I.; Cozens, C.; Abramov, M.; Renders, M.; Zhang, S.; Chaput, J.C.; Wengel, J.; Peak-Chew, S.Y.; McLaughlin, S.H.; et al. Synthetic Genetic Polymers Capable of Heredity and Evolution. Science 2012, 336, 341–344. [Google Scholar] [CrossRef] [Green Version]
  130. Houlihan, G.; Arangundy-Franklin, S.; Porebski, B.T.; Subramanian, N.; Taylor, A.I.; Holliger, P. Discovery and Evolution of RNA and XNA Reverse Transcriptase Function and Fidelity. Nat. Chem. 2020, 12, 683–690. [Google Scholar] [CrossRef] [PubMed]
  131. Paegel, B.M.; Joyce, G.F. Microfluidic Compartmentalized Directed Evolution. Chem. Biol. 2010, 17, 717–724. [Google Scholar] [CrossRef] [Green Version]
  132. Larsen, A.C.; Dunn, M.R.; Hatch, A.; Sau, S.P.; Youngbull, C.; Chaput, J.C. A General Strategy for Expanding Polymerase Function by Droplet Microfluidics. Nat. Commun. 2016, 7, 11235. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Price, A.K.; Paegel, B.M. Discovery in Droplets. Anal. Chem. 2016, 88, 339–353. [Google Scholar] [CrossRef] [Green Version]
  134. Nikoomanzar, A.; Vallejo, D.; Chaput, J.C. Elucidating the Determinants of Polymerase Specificity by Microfluidic-Based Deep Mutational Scanning. ACS Synth. Biol. 2019, 8, 1421–1429. [Google Scholar] [CrossRef]
  135. Taylor, A.I.; Houlihan, G.; Holliger, P. Beyond DNA and RNA: The Expanding Toolbox of Synthetic Genetics. Cold Spring Harbor Perspect. Biol. 2019, 11, a032490. [Google Scholar] [CrossRef] [Green Version]
  136. Ohashi, S.; Hashiya, F.; Abe, H. Variety of Nucleotide Polymerase Mutants Aiming to Synthesize Modified RNA. ChemBioChem 2021, 22, 2398–2406. [Google Scholar] [CrossRef] [PubMed]
  137. Sawai, H.; Ozaki-Nakamura, A.; Mine, M.; Ozaki, H. Synthesis of New Modified DNAs by Hyperthermophilic DNA Polymerase: Substrate and Template Specificity of Functionalized Thymidine Analogues Bearing an sp3-Hybridized Carbon at the C5 Alpha-Position for Several DNA Polymerases. Bioconjugate Chem. 2002, 13, 309–316. [Google Scholar] [CrossRef] [PubMed]
  138. Kuwahara, M.; Nagashima, J.; Hasegawa, M.; Tamura, T.; Kitagata, R.; Hanawa, K.; Hososhima, S.; Kasamatsu, T.; Ozaki, H.; Sawai, H. Systematic Characterization of 2′-Deoxynucleoside-5′-triphosphate Analogs as Substrates for DNA Polymerases by Polymerase Chain Reaction and Kinetic Studies on Enzymatic Production of Modified DNA. Nucleic Acids Res. 2006, 34, 5383–5394. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Mehedi Masud, M.; Ozaki-Nakamura, A.; Kuwahara, M.; Ozaki, H.; Sawai, H. Modified DNA Bearing 5(Methoxycarbonylmethyl)-2′-deoxyuridine: Preparation by PCR with Thermophilic DNA Polymerase and Postsynthetic Derivatization. ChemBioChem 2003, 4, 584–588. [Google Scholar] [CrossRef]
  140. Hottin, A.; Marx, A. Structural Insights into the Processing of Nucleobase-Modified Nucleotides by DNA Polymerases. Accounts Chem. Res. 2016, 49, 418–427. [Google Scholar] [CrossRef]
  141. Jager, S.; Rasched, G.; Kornreich-Leshem, H.; Engeser, M.; Thum, O.; Famulok, M. A Versatile Toolbox for Variable DNA Functionalization at High Density. J. Am. Chem. Soc. 2005, 127, 15071–15082. [Google Scholar] [CrossRef]
  142. Baccaro, A.; Steck, A.L.; Marx, A. Barcoded Nucleotides. Angew. Chem. Int. Edit. 2012, 51, 254–257. [Google Scholar] [CrossRef] [Green Version]
  143. Pochet, S.; Kaminski, P.A.; Van Aerschot, A.; Herdewijn, P.; Marlière, P. Replication of Hexitol Oligonucleotides as a Prelude to the Propagation of a Third Type of Nucleic Acid in Vivo. C. R. Biol. 2003, 326, 1175–1184. [Google Scholar] [CrossRef]
  144. Jackson, L.N.; Chim, N.; Shi, C.H.; Chaput, J.C. Crystal Structures of a Natural DNA Polymerase that Functions as an XNA Reverse Transcriptase. Nucleic Acids Res. 2019, 47, 6973–6983. [Google Scholar] [CrossRef]
  145. Tsai, C.H.; Chen, J.; Szostak, J.W. Enzymatic Synthesis of DNA on Glycerol Nucleic Acid Templates without Stable Duplex Formation between Product and Template. Proc. Natl. Acad. Sci. USA 2007, 104, 14598–14603. [Google Scholar] [CrossRef] [Green Version]
  146. Kempeneers, V.; Renders, M.; Froeyen, M.; Herdewijn, P. Investigation of the DNA-Dependent Cyclohexenyl Nucleic Acid Polymerization and the Cyclohexenyl Nucleic Acid-Dependent DNA Polymerization. Nucleic Acids Res. 2005, 33, 3828–3836. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Chaput, J.C.; Ichida, J.K.; Szostak, J.W. DNA Polymerase-Mediated DNA Synthesis on a TNA Template. J. Am. Chem. Soc. 2003, 125, 856–857. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Peng, C.G.; Damha, M.J. Polymerase-Directed Synthesis of 2′-Deoxy-2′-fluoro-beta-D-arabinonucleic Acids. J. Am. Chem. Soc. 2007, 129, 5310–5311. [Google Scholar] [CrossRef] [PubMed]
  149. Joyce, C.M. Choosing the Right Sugar: How Polymerases Select a Nucleotide Substrate. Proc. Natl. Acad. Sci. USA 1997, 94, 1619–1622. [Google Scholar] [CrossRef] [Green Version]
  150. Anosova, I.; Kowai, E.A.; Dunn, M.R.; Chaput, J.C.; Van Horn, W.D.; Egli, M. The Structural Diversity of Artificial Genetic Polymers. Nucleic Acids Res. 2016, 44, 1007–1021. [Google Scholar] [CrossRef] [Green Version]
  151. Gardner, A.F.; Jackson, K.M.; Boyle, M.M.; Buss, J.A.; Potapov, V.; Gehring, A.M.; Zatopek, K.M.; Corrêa, I.R., Jr.; Ong, J.L.; Jack, W.E. Therminator DNA Polymerase: Modified Nucleotides and Unnatural Substrates. Front. Mol. Biosci. 2019, 6, 28. [Google Scholar] [CrossRef] [Green Version]
  152. Staiger, N.; Marx, A. A DNA Polymerase with Increased Reactivity for Ribonucleotides and C5-Modified Deoxyribonucleotides. ChemBioChem 2010, 11, 1963–1966. [Google Scholar] [CrossRef] [Green Version]
  153. Ichida, J.K.; Horhota, A.; Zou, K.; McLaughlin, L.W.; Szostak, J.W. High Fidelity TNA Synthesis by Therminator Polymerase. Nucleic Acids Res. 2005, 33, 5219–5225. [Google Scholar] [CrossRef] [Green Version]
  154. Fa, M.; Radeghieri, A.; Henry, A.A.; Romesberg, F.E. Expanding the Substrate Repertoire of a DNA Polymerase by Directed Evolution. J. Am. Chem. Soc. 2004, 126, 1748–1754. [Google Scholar] [CrossRef]
  155. Song, P.; Zhang, R.; He, C.; Chen, T. Transcription, Reverse Transcription, and Amplification of Backbone-Modified Nucleic Acids with Laboratory-Evolved Thermophilic DNA Polymerases. Curr. Protoc. 2021, 1, e188. [Google Scholar] [CrossRef]
  156. Chen, T.; Romesberg, F.E. Enzymatic Synthesis, Amplification, and Application of DNA with a Functionalized Backbone. Angew. Chem. Int. Edit. 2017, 56, 14046–14051. [Google Scholar] [CrossRef] [PubMed]
  157. Arangundy-Franklin, S.; Taylor, A.I.; Porebski, B.T.; Genna, V.; Peak-Chew, S.; Vaisman, A.; Woodgate, R.; Orozco, M.; Holliger, P. A Synthetic Genetic Polymer with an Uncharged Backbone Chemistry Based on Alkyl Phosphonate Nucleic Acids. Nat. Chem. 2019, 11, 533–542. [Google Scholar] [CrossRef] [PubMed]
  158. Dunn, M.R.; Otto, C.; Fenton, K.E.; Chaput, J.C. Improving Polymerase Activity with Unnatural Substrates by Sampling Mutations in Homologous Protein Architectures. ACS Chem. Biol. 2016, 11, 1210–1219. [Google Scholar] [CrossRef] [PubMed]
  159. Nikoomanzar, A.; Vallejo, D.; Yik, E.J.; Chaput, J.C. Programmed Allelic Mutagenesis of a DNA Polymerase with Single Amino Acid Resolution. ACS Synth. Biol. 2020, 9, 1873–1881. [Google Scholar] [CrossRef] [PubMed]
  160. Liu, C.; Cozens, C.; Jaziri, F.; Rozenski, J.; Marechal, A.; Dumbre, S.; Pezo, V.; Marlière, P.; Pinheiro, V.B.; Groaz, E.; et al. Phosphonomethyl Oligonucleotides as Backbone-Modified Artificial Genetic Polymers. J. Am. Chem. Soc. 2018, 140, 6690–6699. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  161. Hoshino, H.; Kasahara, Y.; Kuwahara, M.; Obika, S. DNA Polymerase Variants with High Processivity and Accuracy for Encoding and Decoding Locked Nucleic Acid Sequences. J. Am. Chem. Soc. 2020, 142, 21530–21537. [Google Scholar] [CrossRef] [PubMed]
  162. Medina, E.; Yik, E.J.; Herdewijn, P.; Chaput, J.C. Functional Comparison of Laboratory-Evolved XNA Polymerases for Synthetic Biology. ACS Synth. Biol. 2021, 10, 1429–1437. [Google Scholar] [CrossRef]
  163. Ghadessy, F.J.; Ramsay, N.; Boudsocq, F.; Loakes, D.; Brown, A.; Iwai, S.; Vaisman, A.; Woodgate, R.; Holliger, P. Generic Expansion of the Substrate Spectrum of a DNA Polymerase by Directed Evolution. Nat. Biotechnol. 2004, 22, 755–759. [Google Scholar] [CrossRef]
  164. Cozens, C.; Pinheiro, V.B.; Vaisman, A.; Woodgate, R.; Holliger, P. A Short Adaptive Path from DNA to RNA Polymerases. Proc. Natl. Acad. Sci. USA 2012, 109, 8067–8072. [Google Scholar] [CrossRef]
  165. Cozens, C.; Mutschler, H.; Nelson, G.M.; Houlihan, G.; Taylor, A.I.; Holliger, P. Enzymatic Synthesis of Nucleic Acids with Defined Regioisomeric 2′-5′ Linkages. Angew. Chem. Int. Edit. 2015, 54, 15570–15573. [Google Scholar] [CrossRef] [Green Version]
  166. Freund, N.; Taylor, A.I.; Arangundy-Franklin, S.; Subramanian, N.; Peak-Chew, S.Y.; Whitaker, A.M.; Freudenthal, B.D.; Abramov, M.; Herdewijn, P.; Holliger, P. A Two-Residue Nascent-Strand Steric Gate Controls Synthesis of 2′-O-methyl- and 2′-O-(2-methoxyethyl)-RNA. Nat. Chem. 2022. [Google Scholar] [CrossRef] [PubMed]
  167. Li, Q.; Maola, V.A.; Chim, N.; Hussain, J.; Lozoya-Colinas, A.; Chaput, J.C. Synthesis and Polymerase Recognition of Threose Nucleic Acid Triphosphates Equipped with Diverse Chemical Functionalities. J. Am. Chem. Soc. 2021, 143, 17761–17768. [Google Scholar] [CrossRef] [PubMed]
  168. Brown, J.A.; Suo, Z. Unlocking the Sugar “Steric Gate” of DNA Polymerases. Biochemistry 2011, 50, 1135–1142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Vichier-Guerre, S.; Ferris, S.; Auberger, N.; Mahiddine, K.; Jestin, J.L. A Population of Thermostable Reverse Transcriptases Evolved from Thermus aquaticus DNA Polymerase I by Phage Display. Angew. Chem. Int. Edit. 2006, 45, 6133–6137. [Google Scholar] [CrossRef] [PubMed]
  170. Blatter, N.; Bergen, K.; Nolte, O.; Welte, W.; Diederichs, K.; Mayer, J.; Wieland, M.; Marx, A. Structure and Function of an RNA-Reading Thermostable DNA Polymerase. Angew. Chem. Int. Edit. 2013, 52, 11935–11939. [Google Scholar] [CrossRef] [Green Version]
  171. Aschenbrenner, J.; Marx, A. Direct and Site-Specific Quantification of RNA 2′-O-methylation by PCR with an Engineered DNA Polymerase. Nucleic Acids Res. 2016, 44, 3495–3502. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Fogg, M.J.; Pearl, L.H.; Connolly, B.A. Structural Basis for Uracil Recognition by Archaeal Family B DNA Polymerases. Nat. Struct. Biol. 2002, 9, 922–927. [Google Scholar] [CrossRef]
  173. Chim, N.; Shi, C.; Sau, S.P.; Nikoomanzar, A.; Chaput, J.C. Structural Basis for TNA Synthesis by an Engineered TNA Polymerase. Nat. Commun. 2017, 8, 1810. [Google Scholar] [CrossRef] [Green Version]
  174. Horhota, A.; Zou, K.; Ichida, J.K.; Yu, B.; McLaughlin, L.W.; Szostak, J.W.; Chaput, J.C. Kinetic Analysis of an Efficient DNA-Dependent TNA Polymerase. J. Am. Chem. Soc. 2005, 127, 7427–7434. [Google Scholar] [CrossRef]
  175. Gardner, A.F.; Jack, W.E. Determinants of Nucleotide Sugar Recognition in an Archaeon DNA Polymerase. Nucleic Acids Res. 1999, 27, 2545–2553. [Google Scholar] [CrossRef] [Green Version]
  176. Duffy, K.; Arangundy-Franklin, S.; Holliger, P. Modified Nucleic Acids: Replication, Evolution, and Next-Generation Therapeutics. BMC Biol. 2020, 18, 112. [Google Scholar] [CrossRef] [PubMed]
  177. Chan, J.H.; Lim, S.; Wong, W.S. Antisense Oligonucleotides: From Design to Therapeutic Application. Clin. Exp. Pharmacol. Physiol. 2006, 33, 533–540. [Google Scholar] [CrossRef] [PubMed]
  178. Eckstein, F. Phosphorothioates, Essential Components of Therapeutic Oligonucleotides. Nucl. Acid Ther. 2014, 24, 374–387. [Google Scholar] [CrossRef] [PubMed]
  179. Perry, C.M.; Balfour, J.A. Fomivirsen. Drugs 1999, 57, 375–380; discussion 381. [Google Scholar] [CrossRef]
  180. Liang, X.H.; Sun, H.; Shen, W.; Crooke, S.T. Identification and Characterization of Intracellular Proteins that Bind Oligonucleotides with Phosphorothioate Linkages. Nucleic Acids Res. 2015, 43, 2927–2945. [Google Scholar] [CrossRef] [Green Version]
  181. Stein, C.A.; Subasinghe, C.; Shinozuka, K.; Cohen, J.S. Physicochemical Properties of Phosphorothioate Oligodeoxynucleotides. Nucleic Acids Res. 1988, 16, 3209–3221. [Google Scholar] [CrossRef]
  182. Kurreck, J. Antisense Technologies. Improvement through Novel Chemical Modifications. Eur. J. Biochem. 2003, 270, 1628–1644. [Google Scholar] [CrossRef]
  183. Bennett, C.F.; Swayze, E.E. RNA Targeting Therapeutics: Molecular Mechanisms of Antisense Oligonucleotides as a Therapeutic Platform. Annu. Rev. Pharmacol. Toxicol. 2010, 50, 259–293. [Google Scholar] [CrossRef]
  184. Geary, R.S.; Baker, B.F.; Crooke, S.T. Clinical and Preclinical Pharmacokinetics and Pharmacodynamics of Mipomersen (Kynamro®): A Second-Generation Antisense Oligonucleotide Inhibitor of Apolipoprotein B. Clin. Pharmacokinet. 2015, 54, 133–146. [Google Scholar] [CrossRef]
  185. Lok, C.N.; Viazovkina, E.; Min, K.L.; Nagy, E.; Wilds, C.J.; Damha, M.J.; Parniak, M.A. Potent Gene-Specific Inhibitory Properties of Mixed-Backbone Antisense Oligonucleotides Comprised of 2′-Deoxy-2′-fluoro-d-arabinose and 2′-Deoxyribose Nucleotides. Biochemistry 2002, 41, 3457–3467. [Google Scholar] [CrossRef]
  186. Damha, M.J.; Wilds, C.J.; Noronha, A.; Brukner, I.; Borkow, G.; Arion, D.; Parniak, M.A. Hybrids of RNA and Arabinonucleic Acids (ANA and 2′ F-ANA) Are Substrates of Ribonuclease h. J. Am. Chem. Soc. 1998, 120, 12976–12977. [Google Scholar] [CrossRef]
  187. Fluiter, K.; Frieden, M.; Vreijling, J.; Rosenbohm, C.; De Wissel, M.B.; Christensen, S.M.; Koch, T.; Ørum, H.; Baas, F. On the in Vitro and in Vivo Properties of Four Locked Nucleic Acid Nucleotides Incorporated into an Anti-H-Ras Antisense Oligonucleotide. ChemBioChem 2005, 6, 1104–1109. [Google Scholar] [CrossRef] [PubMed]
  188. Kaur, H.; Babu, B.R.; Maiti, S. Perspectives on Chemistry and Therapeutic Applications of Locked Nucleic Acid (LNA). Chem. Rev. 2007, 107, 4672–4697. [Google Scholar] [CrossRef] [PubMed]
  189. Walter, N.G.; Engelke, D.R. Ribozymes: Catalytic RNAs That Cut Things, Make Things, and Do Odd and Useful Jobs. Biologist 2002, 49, 199–203. [Google Scholar] [PubMed]
  190. Micura, R.; Höbartner, C. Fundamental Studies of Functional Nucleic Acids: Aptamers, Riboswitches, Ribozymes and DNAzymes. Chem. Soc. Rev. 2020, 49, 7331–7353. [Google Scholar] [CrossRef] [PubMed]
  191. Ma, L.; Liu, J. Catalytic Nucleic Acids: Biochemistry, Chemical Biology, Biosensors, and Nanotechnology. iScience 2020, 23, 100815. [Google Scholar] [CrossRef] [Green Version]
  192. Taylor, A.I.; Pinheiro, V.B.; Smola, M.J.; Morgunov, A.S.; Peak-Chew, S.; Cozens, C.; Weeks, K.M.; Herdewijn, P.; Holliger, P. Catalysts from Synthetic Genetic Polymers. Nature 2015, 518, 427–430. [Google Scholar] [CrossRef] [Green Version]
  193. Wang, Y.; Ngor, A.K.; Nikoomanzar, A.; Chaput, J.C. Evolution of a General RNA-Cleaving FANA Enzyme. Nat. Commun. 2018, 9, 5067. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Wang, Y.; Nguyen, K.; Spitale, R.C.; Chaput, J.C. A Biologically Stable DNAzyme That Efficiently Silences Gene Expression in Cells. Nat. Chem. 2021, 13, 319–326. [Google Scholar] [CrossRef]
  195. Wilds, C.J.; Damha, M.J. 2′-Deoxy-2′-fluoro-beta-D-arabinonucleosides and Oligonucleotides (2′F-ANA): Synthesis and Physicochemical Studies. Nucleic Acids Res. 2000, 28, 3625–3635. [Google Scholar] [CrossRef]
  196. Wang, Y.; Wang, Y.; Song, D.; Sun, X.; Zhang, Z.; Li, X.; Li, Z.; Yu, H. A Threose Nucleic Acid Enzyme with RNA Ligase Activity. J. Am. Chem. Soc. 2021, 143, 8154–8163. [Google Scholar] [CrossRef] [PubMed]
  197. Wang, Y.; Wang, Y.; Song, D.F.; Sun, X.; Li, Z.; Chen, J.Y.; Yu, H. An RNA-Cleaving Threose Nucleic Acid Enzyme Capable of Single Point Mutation Discrimination. Nat. Chem. 2022, 14, 350–359. [Google Scholar] [CrossRef] [PubMed]
  198. Ran, F.A.; Hsu, P.D.; Wright, J.; Agarwala, V.; Scott, D.A.; Zhang, F. Genome Engineering Using the CRISPR-Cas9 System. Nat. Protoc. 2013, 8, 2281–2308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Rahdar, M.; McMahon, M.A.; Prakash, T.P.; Swayze, E.E.; Bennett, C.F.; Cleveland, D.W. Synthetic CRISPR RNA-Cas9–Guided Genome Editing in Human Cells. Proc. Natl. Acad. Sci. USA 2015, 112, E7110–E7117. [Google Scholar] [CrossRef] [Green Version]
  200. Ryan, D.E.; Taussig, D.; Steinfeld, I.; Phadnis, S.M.; Lunstad, B.D.; Singh, M.; Vuong, X.; Okochi, K.D.; McCaffrey, R.; Olesiak, M.; et al. Improving CRISPR–Cas Specificity with Chemical Modifications in Single-Guide RNAs. Nucleic Acids Res. 2018, 46, 792–803. [Google Scholar] [CrossRef]
  201. Mir, A.; Alterman, J.F.; Hassler, M.R.; Debacker, A.J.; Hudgens, E.; Echeverria, D.; Brodsky, M.H.; Khvorova, A.; Watts, J.K.; Sontheimer, E.J. Heavily and Fully Modified RNAs Guide Efficient SpyCas9-Mediated Genome Editing. Nat. Commun. 2018, 9, 2641. [Google Scholar] [CrossRef] [Green Version]
  202. Cromwell, C.R.; Sung, K.; Park, J.; Krysler, A.R.; Jovel, J.; Kim, S.K.; Hubbard, B.P. Incorporation of Bridged Nucleic Acids into CRISPR RNAs Improves Cas9 Endonuclease Specificity. Nat. Commun. 2018, 9, 1448. [Google Scholar] [CrossRef] [Green Version]
  203. Kong, H.Y.; Byun, J. Nucleic Acid Aptamers: New Methods for Selection, Stabilization, and Application in Biomedical Science. Biomol. Ther. 2013, 21, 423–434. [Google Scholar] [CrossRef] [Green Version]
  204. Ng, E.W.; Shima, D.T.; Calias, P.; Cunningham, E.T., Jr.; Guyer, D.R.; Adamis, A.P. Pegaptanib, a Targeted Anti-VEGF Aptamer for Ocular Vascular Disease. Nat. Rev. Drug Discov. 2006, 5, 123–132. [Google Scholar] [CrossRef]
  205. Liu, Z.; Chen, T.; Romesberg, F.E. Evolved Polymerases Facilitate Selection of Fully 2′-OMe-Modified Aptamers. Chem. Sci. 2017, 8, 8179–8182. [Google Scholar] [CrossRef] [Green Version]
  206. Rangel, A.E.; Chen, Z.; Ayele, T.M.; Heemstra, J.M. In Vitro Selection of an XNA Aptamer Capable of Small-Molecule Recognition. Nucleic Acids Res. 2018, 46, 8057–8068. [Google Scholar] [CrossRef] [PubMed]
  207. Rose, K.M.; Alves Ferreira-Bravo, I.; Li, M.; Craigie, R.; Ditzler, M.A.; Holliger, P.; DeStefano, J.J. Selection of 2′-Deoxy-2′-Fluoroarabino Nucleic Acid (FANA) Aptamers that Bind HIV-1 Integrase with Picomolar Affinity. ACS Chem. Biol. 2019, 14, 2166–2175. [Google Scholar] [CrossRef] [PubMed]
  208. Alves Ferreira-Bravo, I.; Cozens, C.; Holliger, P.; DeStefano, J.J. Selection of 2′-Deoxy-2′-Fluoroarabinonucleotide (FANA) Aptamers That Bind HIV-1 Reverse Transcriptase with Picomolar Affinity. Nucleic Acids Res. 2015, 43, 9587–9599. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Inoue, N.; Shionoya, A.; Minakawa, N.; Kawakami, A.; Ogawa, N.; Matsuda, A. Amplification of 4′-ThioDNA in the Presence of 4′-Thio-dTTP and 4′-Thio-dCTP, and 4′-Thio-DNA-Directed Transcription in Vitro and in Mammalian Cells. J. Am. Chem. Soc. 2007, 129, 15424–15425. [Google Scholar] [CrossRef] [PubMed]
  210. Pezo, V.; Liu, F.W.; Abramov, M.; Froeyen, M.; Herdewijn, P.; Marlière, P. Binary Genetic Cassettes for Selecting XNA-Templated DNA Synthesis in Vivo. Angew. Chem. Int. Edit. 2013, 125, 8297–8301. [Google Scholar] [CrossRef]
  211. Nguyen, H.; Abramov, M.; Eremeeva, E.; Herdewijn, P. In Vivo Expression of Genetic Information from Phosphoramidate–DNA. ChemBioChem 2020, 21, 272–278. [Google Scholar] [CrossRef]
  212. Kukwikila, M.; Gale, N.; El-Sagheer, A.H.; Brown, T.; Tavassoli, A. Assembly of a Biocompatible Triazole-Linked Gene by One-Pot Click-DNA Ligation. Nat. Chem. 2017, 9, 1089–1098. [Google Scholar] [CrossRef]
  213. Kim, K.R.; Kim, H.Y.; Lee, Y.D.; Ha, J.S.; Kang, J.H.; Jeong, H.; Bang, D.; Ko, Y.T.; Kim, S.; Lee, H.; et al. Self-Assembled Mirror DNA Nanostructures for Tumor-Specific Delivery of Anticancer Drugs. J. Control. Release 2016, 243, 121–131. [Google Scholar] [CrossRef]
  214. Gamberi, G.; Morandi, L.; Benini, S.; Resca, A.; Cocchi, S.; Magagnoli, G.; Donati, D.M.; Righi, A.; Gambarotti, M. Detection of H3F3A p. G35W and p. G35R in Giant Cell Tumor of Bone by Allele Specific Locked Nucleic Acid Quantitative PCR (ASLNAqPCR). Pathol. Res. Pract. 2018, 214, 89–94. [Google Scholar] [CrossRef]
  215. Chaput, J.C. Redesigning the Genetic Polymers of Life. Accounts Chem. Res. 2021, 54, 1056–1065. [Google Scholar] [CrossRef]
  216. Mana, T.; Bhattacharya, B.; Lahiri, H.; Mukhopadhyay, R. XNAs: A Troubleshooter for Nucleic acid Sensing. Acs Omega 2022, 7, 15296–15307. [Google Scholar] [CrossRef] [PubMed]
  217. Freund, N.; Fürst, M.J.L.J.; Holliger, P. New Chemistries and Enzymes for Synthetic Genetics. Curr. Opin. Biotechnol. 2022, 74, 129–136. [Google Scholar] [CrossRef]
  218. Taylor, A.I.; Beuron, F.; Peak-Chew, S.Y.; Morris, E.P.; Herdewijn, P.; Holliger, P. Nanostructures from Synthetic Genetic Polymers. ChemBioChem 2016, 17, 1107–1110. [Google Scholar] [CrossRef]
  219. Pavlov, A.R.; Pavlova, N.V.; Kozyavkin, S.A.; Slesarev, A.I. Recent Developments in the Optimization of Thermostable DNA Polymerases for Efficient Applications. Trends Biotechnol. 2004, 22, 253–260. [Google Scholar] [CrossRef] [PubMed]
  220. Pinheiro, V.B.; Holliger, P. The XNA World: Progress towards Replication and Evolution of Synthetic Genetic Polymers. Curr. Opin. Chem. Biol. 2012, 16, 245–252. [Google Scholar] [CrossRef] [PubMed]
  221. Pinheiro, V.B.; Holliger, P. Towards XNA Nanotechnology: New Materials from Synthetic Genetic Polymers. Trends Biotechnol. 2014, 32, 321–328. [Google Scholar] [CrossRef] [Green Version]
  222. Malik, T.N.; Chaput, J.C. XNA Enzymes by Evolution and Design. Curr. Res. Chem. Biol. 2021, 1, 100012. [Google Scholar] [CrossRef]
Figure 1. Phylogenetic tree of representative thermophilic DNAPs. Green: family A DNAPs; pale yellow: family B DNAPs.
Figure 1. Phylogenetic tree of representative thermophilic DNAPs. Green: family A DNAPs; pale yellow: family B DNAPs.
Ijms 23 14969 g001
Figure 2. Expansion of the central dogma with XNAs and XNAPs. Green arrows: replication; blue arrows: transcription or reverse transcription.
Figure 2. Expansion of the central dogma with XNAs and XNAPs. Green arrows: replication; blue arrows: transcription or reverse transcription.
Ijms 23 14969 g002
Figure 3. Representative methods employed for the generation of XNAPs. (a) Methods for the construction of polymerase libraries. I: error-prone PCR; II: site-directed saturation mutagenesis; III: gene shuffling by StEP; IV: gene shuffling by DNase I digestion and PCR reassembly; (b) methods for the selection of polymerase mutants. I: phage display; II: CSR; III: CST.
Figure 3. Representative methods employed for the generation of XNAPs. (a) Methods for the construction of polymerase libraries. I: error-prone PCR; II: site-directed saturation mutagenesis; III: gene shuffling by StEP; IV: gene shuffling by DNase I digestion and PCR reassembly; (b) methods for the selection of polymerase mutants. I: phage display; II: CSR; III: CST.
Ijms 23 14969 g003
Figure 4. Summary of the relationships and mutations of some representative engineered XNAPs. The mutated amino acids in the engineered XNAPs are indicated in red.
Figure 4. Summary of the relationships and mutations of some representative engineered XNAPs. The mutated amino acids in the engineered XNAPs are indicated in red.
Ijms 23 14969 g004
Figure 5. Distribution of the mutation sites in the engineered XNAPs. The mutation sites in engineered (a) Taq DNAP (green, PDB: 1TAU); (b) Tgo DNAP (cyan, PDB: 7B07); (c) KOD DNAP (yellow, PDB: 4K8Z); and (d) 9°N DNAP (red, PDB: 6ISF). The DNA templates and DNA primers are shown in blue and orange, respectively.
Figure 5. Distribution of the mutation sites in the engineered XNAPs. The mutation sites in engineered (a) Taq DNAP (green, PDB: 1TAU); (b) Tgo DNAP (cyan, PDB: 7B07); (c) KOD DNAP (yellow, PDB: 4K8Z); and (d) 9°N DNAP (red, PDB: 6ISF). The DNA templates and DNA primers are shown in blue and orange, respectively.
Ijms 23 14969 g005
Figure 6. Application of XNAs and XNAPs. Modifications endow XNAs with expanded structural and functional diversities, and XNAPs further broaden the application scope of XNAs. (a) XNAzymes; (b) modified antisense oligonucleotides; (c) modified guide RNAs for CRISPR/Cas9 system; (d) XNA aptamers; (e) genetic information storage in living organisms; (f) XNA materials.
Figure 6. Application of XNAs and XNAPs. Modifications endow XNAs with expanded structural and functional diversities, and XNAPs further broaden the application scope of XNAs. (a) XNAzymes; (b) modified antisense oligonucleotides; (c) modified guide RNAs for CRISPR/Cas9 system; (d) XNA aptamers; (e) genetic information storage in living organisms; (f) XNA materials.
Ijms 23 14969 g006
Table 1. Representative thermophilic and hyperthermophilic DNAPs.
Table 1. Representative thermophilic and hyperthermophilic DNAPs.
FamilyDNAPSourcePropertiesRef.
5′-3′ Exo3′-5′ ExoError RateHalf-Life Time
ATaqThermus aquaticusYesNo1.2 × 10−5–3.3 × 10−697.5 °C/9 min [25,26]
TfiThermus filiformisYesNo/94 °C/40 min[27]
TthThermus thermophilusYesNo/94 °C/20 min[10,28]
TflThermus flavusYesNo/95 °C/40 min[9,29]
TcaThermus caldophilusYesNo/95 °C/70 min[30]
TsK1Thermus scotoductusYesNo/95 °C/15 min[31]
BstBacillus stearothermophilusYesNo//[32]
BcaBacillus caldotenaxYesNo//[33]
BcavBacillus caldoveloxYesNo//[34]
BsmBacillus smithiiYesNo//[35]
GssGeobacillus sp. 777YesNo//[36]
TmaThermotoga maritimaYesYes//[37]
TneThermotoga neapolitanaYesYes3.4 × 10−5/[38,39]
AaeAquifex aeolicusNoYes/75 °C/6 h
85 °C/1.7 h
[40]
BTliThermococcus litoralisNoYes2.8 × 10−6100 °C/2 h[41]
KODThermococcus kodakaraensisNoYes2.6 × 10−695 °C/12 h[42]
9°NThermococcus sp. 9°N-7NoYes//[43]
TgoThermococcus gorganariusNoYes3.3–2.2 × 10−6/[44]
TfuThermococcus fumicolansNoYes5.3–0.9 × 10−5100 °C/2 h [45]
TNA1Thermococcus sp. NA1NoYes2.2 × 10−495 °C/12.5 h
100 °C/3.5 h
[46]
TpeThermococcus peptonophilusNoYes3.37 × 10−690 °C/4 h[47]
TziThermococcus zilligiiNoYes2 × 10−6/[48]
TwaThermococcus waiotapuensisNoYes7.4 × 10−699 °C/4 h[49]
PfuPyrococcus furiosusNoYes1.3 × 10−6/[50]
PstPyrococcus GB-DNoYes2.7 × 10−695 °C/23 h[50,51]
PabPyrococcus abyssiNoYes0.66–1.39 × 10−6100 °C/5 h[52]
PwoPyrococcus woeseiNoYes/95 °C/8 h[10,53]
Table 2. Summary of engineered thermophilic and hyperthermophilic XNAPs.
Table 2. Summary of engineered thermophilic and hyperthermophilic XNAPs.
Parental DNAPMutantMethod Employed for EngineeringMutation SitesUnnatural ActivityRef.
TaqAA40spCSRE602V, A608V, I614M, E615GSynthesis of 2′-F, 2′-N3 and 2′-OMe-RNA[125]
SFM19Phage displayI614E, E615GSynthesis of 2′-OMe-modified RNA[154]
SFM4-3Phage displayI614E, E615G, V518A, N583S, D655N, E681K, E742Q, M747RSynthesis or amplification of 2′-OMe, 2′-F, 2′-Az, 2′-Cl, 2′-Am-modified DNA/RNA and ANA[120]
SFM4-6Phage displayI614E, E615G, D655N, L657M, E681K, E742N, M747RSynthesis of 2′-F-DNA and 2′-OMe-RNA[120]
SFM4-9Phage displayI614E, E615G, N415Y, V518A, D655N, L657M, E681V, E742N, M747RSynthesis of DNA from a 2′-F-DNA or 2′-OMe-RNA template[120]
M1CSRG84A, D144G, K314R, E520G, F598L, A608V, E742GPCR of phosphorothioate or fluorescent dye-modified DNA[163]
M4CSRD58G, R74P, A109T, L245R, R343G, G370D, E520G, N583S, E694K, A743PPCR of phosphorothioate or fluorescent dye-modified DNA[163]
TgoTgo-RISDM *D141A, E143A, A485R, E664ISynthesis of TNA[158]
Tgo TGKSDM *TgoT: Y409G, E664KSynthesis of pseudouridine-, 5-methyl-C-, 2′-F-, 2′-Az-modified RNAs, FANA, ANA, HNA and TNA[162,164]
Tgo TGLLKSDM *TgoT: Y409G, I521L, F545L, E664KSynthesis of 3′-deoxy- or 3′-O-methyl-modified RNA[165]
RT521CSTTgoT: E429G, I521L, K726RSynthesis of DNA from an HNA, ANA, FANA or tPhoNA template[129]
RT521KCSTRT521: F445L, E664KSynthesis of DNA from an LNA or CeNA template[129]
RT-TKKCBLRT521K: I114T, S383K, N735KSynthesis of DNA from a 2′-OMe-RNA or AtNA template[130]
RT-C8CBLRT-TKK: F493V, Y496N, Y497L, Y499A, A500Q, K501HSynthesis of DNA from a 2′-OMe-RNA, HNA, AtNA, 2′-MOE-RNA or PS 2′-MOE-RNA template[130]
RT-C8exo+SDM *RT-TKK: A141D, A143E, F493V, Y496N, Y497L, Y499A, A500Q, K501H Synthesis of DNA from a 2′-OMe-RNA, HNA, AtNA, 2′-MOE-RNA or PS 2′-MOE-RNA template[130]
RT-H4CBLRT-TKK: F493V, Y496H, Y497M, Y499F, A500E, K501NSynthesis of DNA from an HNA template[130]
RT-H4exo+SDM *RT-TKK: A141D, A143E, F493V, Y496H, Y497M, Y499F, A500E, K501NSynthesis of DNA from an HNA template[130]
RT-TRCBLRT521K: P410T, S411RSynthesis of DNA from a 2′-OMe-RNA, HNA, AtNAs, 2′-MOE-RNA or PS 2′-MOE-RNA template with enhanced fidelity[130]
PolC7CSTTgoT: K659Q, V661A, E664Q, Q665P, D669A, K671Q, T676K, R709KSynthesis of CeNA and LNA[129]
PolD4KCSTTgoT: L403P, P657T, E658Q, K659H, Y663H, E664K, D669A, K671N, T676ISynthesis of FANA, ANA, TNA, HNA, and PMT[129,162]
Pol6G12CSTTgoT: V589A, E609K, I610M, K659Q, E664Q, Q665P, R668K, D669Q, K671H, K674R, T676R, A681S, L704P, E730GSynthesis of HNA and FANA[129,162]
6G12-I521LSDM *Pol6G12: I521LSynthesis of HNA and FANA[162]
Tgo EPFLHSDM *V93Q, D141A, E143A, H147E, L403P, L408F, A485L, I521L, E664HSynthesis of PMT, ANA, TNA, FANA and tPhoNA[160,162]
2MSDM *TGLLK: T541G, K592ASynthesis of 2′-MOE-RNA and 2′-OMe-RNA[166]
3MSDM *TGLLK: T541G, K592A, K664RSynthesis of 2′-MOE-RNA and 2′-OMe-RNA[166]
KODKOD DGLNKSDM *N210D, Y409G, A485L, D614N, E664KSynthesis of 2′-OMe-RNA and LNA[161]
KOD DLKSDM *N210D, A485L, E664KSynthesis of DNA from an LNA template[161]
Kod RISDM *D141A, E143A, A485R, E664ISynthesis of TNA[158]
Kod RSDrOPSD141A, E143A, A485R, N491SSynthesis of TNA[134]
Kod QSDrOPSD141A, E143A, L489Q, N491SSynthesis of TNA[134]
Kod RSGADrOPSD141A, E143A, A485R, N491S, R606G, T723ASynthesis of FANA, ANA, HNA, TNA, C5-modified TNA, and PMT[159,162,167]
KOD RTXRT-CSRF38L, R97M, K118I, M137L, R381H, Y384H, V389I, K466R, Y493L, T514I, I521L, F587L, E664K, G711V, N735K, W768RSynthesis of DNA from a 2′-OMe-RNA template[126]
KOD RTX-Ome v6RT-CSRRTX: A40V, E251K, S340P, G350V, V353L, H381R, H384Y, K468N, I488L, G498A, K664RSynthesis of DNA from a 2′-OMe-RNA template[127]
KOD RT521KSDM *V93E, D141A, E143A, A485L, I521L, E664KSynthesis of DNA from a tPhoNA template[160]
9°N9°N
Therminator
SDM *D141A, E143A, A485LSynthesis of TNA[153]
9n-YRIDrOPSD141A, E143A, A485R,
E664I
Synthesis of TNA[132]
9n-
NVA
DrOPSD141A, E143A, V409N, A485V, E664A, D432G, V636ASynthesis of TNA[132]
Deep VentDeep Vent RISDM *D141A, E143A, A485R, E664ISynthesis of TNA[158]
* SDM: site-directed mutagenesis (including site-directed saturation mutagenesis).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, G.; Du, Y.; Ma, X.; Ye, F.; Qin, Y.; Wang, Y.; Xiang, Y.; Tao, R.; Chen, T. Thermophilic Nucleic Acid Polymerases and Their Application in Xenobiology. Int. J. Mol. Sci. 2022, 23, 14969. https://doi.org/10.3390/ijms232314969

AMA Style

Wang G, Du Y, Ma X, Ye F, Qin Y, Wang Y, Xiang Y, Tao R, Chen T. Thermophilic Nucleic Acid Polymerases and Their Application in Xenobiology. International Journal of Molecular Sciences. 2022; 23(23):14969. https://doi.org/10.3390/ijms232314969

Chicago/Turabian Style

Wang, Guangyuan, Yuhui Du, Xingyun Ma, Fangkai Ye, Yanjia Qin, Yangming Wang, Yuming Xiang, Rui Tao, and Tingjian Chen. 2022. "Thermophilic Nucleic Acid Polymerases and Their Application in Xenobiology" International Journal of Molecular Sciences 23, no. 23: 14969. https://doi.org/10.3390/ijms232314969

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop