30.01.2013 Views

Givaudan-Roure Lecture - Association for Chemoreception Sciences

Givaudan-Roure Lecture - Association for Chemoreception Sciences

Givaudan-Roure Lecture - Association for Chemoreception Sciences

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

1 <strong>Givaudan</strong>-<strong>Roure</strong> <strong>Lecture</strong> [ ] <strong>Givaudan</strong>-<strong>Roure</strong> <strong>Lecture</strong><br />

WHY NEW NEURONS IN THE ADULT OLFACTORY BULB<br />

Alvarez-Buylla A. 1 1Department of Neurological Surgery and Program<br />

in Developmental and Stem Cell Biology, University of Cali<strong>for</strong>nia at<br />

San Francisco, San Francisco, CA<br />

New neurons continue to be <strong>for</strong>med in adult brains. These cells<br />

incorporate into mature circuitry raising questions on how these cells<br />

are <strong>for</strong>med and what is their functional contribution. Adult<br />

neurogenesis is most prominent in the subventricular zone (SVZ) on the<br />

lateral walls of the lateral ventricles. This process has now been<br />

described in many species including primates. Young neurons born in<br />

the SVZ engage in an extraordinary journey, along the walls of the<br />

lateral ventricle and the rostral migratory stream that leads into the<br />

olfactory bulb, where these cells differentiate into granule and<br />

periglomerular interneurons. These cells must be continually replaced<br />

in adult life, as their total number remains unchanged in adults. I will<br />

review recent advances on the origin, mechanism of migration and<br />

integration of these new neurons in adult brain. I will describe the<br />

different stages in the maturation and present electrophysiological<br />

evidence demonstrating that these neurons become functionally<br />

incorporated in adult olfactory bulb circuits. Surprisingly, soon after<br />

full maturation and incorporation many of the new neurons die in a<br />

process that appears to be dependent on olfactory activity. A computer<br />

model of a neural network that approximates the circuitry of the<br />

olfactory bulb predicts that replacement of neurons through an activitydependent<br />

mechanism greatly improves odor discrimination. A<br />

similar principle could apply to other circuits that receive new neuron in<br />

adults.<br />

2 Slide [ ] Olfactory Bulb Physiology<br />

UNDERSTANDING THE ROLE OF GAP JUNCTIONS IN<br />

OLFACTORY FUNCTION<br />

Zhang C. 1, Restrepo D. 1 1Department of Cellular and Developmental<br />

Biology, Neuroscience Program and the Rocky Mountain Smell and<br />

Taste Center, University of Colorado Health <strong>Sciences</strong> Center, Denver,<br />

CO<br />

Previously, we reported that we had generated OlfDNCX mice that<br />

express a dominant negative connexin 43 protein (Cx43/&beta-gal) in<br />

mature olfactory receptor neurons (ORNs) to inactivate Cx43 gap<br />

junctions (Chem. Sense 27:664). The use of olfactory marker protein<br />

promoter to drive the expression of Cx43/&beta-gal ensures the<br />

dysfunction of gap junctions occurs only in mature ORNs, without<br />

affecting the function of gap junctions in other epithelial cells and the<br />

maturation of olfactory epithelial cells. Using electro-olfactogram<br />

(EOG) recordings, we found that 500 &muM IBMX, a<br />

phosphodiesterase inhibitor that elicits responses by increasing<br />

intracellular cAMP, consistently induced large olfactory responses in<br />

various location in the olfactory epithelium and that the magnitude did<br />

not differ between OlfDNCX and controls. This suggests that<br />

Cx43/&beta-gal expression in the olfactory epithelium in OlfDNCX did<br />

not result in a gross interference of signal transduction machinery in<br />

ORNs. We decided to conduct a systematic study to understand how<br />

dysfunction of gap junctions in mature ORNs will affect responses to<br />

odors at the peripheral level, odor maps in the olfactory bulb and<br />

olfactory sensation. We found that OlfDNCX had reduced EOG<br />

responses to select odors, including octaldehyde, which led to “simpler”<br />

odor maps in the olfactory bulb as compared to controls. Gap junctions<br />

in the olfactory epithelium are likely to play a role in odor detection and<br />

discrimination.<br />

This work was supported by grants DC04952, DC00566 and<br />

DC04657 from the NIDCD.<br />

1<br />

3 Slide [ ] Olfactory Bulb Physiology<br />

REAL-TIME IMAGING OF ODORANT-STIMULATED NITRIC<br />

OXIDE PRODUCTION IN THE ANTENNAL LOBE OF<br />

MANDUCA SEXTA<br />

Nighorn A. 1, Carlsson M. 2, Hansson B. 3, Collmann C. 1 1Neurobiology,<br />

University of Arizona, Tucson, AZ; 2Swedish Agricultural University,<br />

Alnarp, Sweden; 3Swedish University of Agricultural <strong>Sciences</strong>, Alnarp,<br />

Sweden<br />

In the antennal lobe (AL) of the moth, Manduca sexta, nitric oxide<br />

synthase is expressed exclusively in the axons of most or all olfactory<br />

receptor neurons (ORNs), and soluble guanylyl cyclase is expressed in<br />

subsets of AL neurons, which suggests that the gaseous signaling<br />

molecule, nitric oxide (NO), has a role in processing olfactory<br />

in<strong>for</strong>mation. To determine if NO is produced by odorant-stimulated<br />

ORNs, we used real-time imaging with a fluorescent NO marker, DAF-<br />

FM diacetate, to record odorant-stimulated NO production in AL<br />

glomeruli. Odorants caused a reproducible, spatially-distinct, and<br />

concentration-dependent production of NO. Different plant odorants<br />

evoked spatially-distinct patterns of NO production in sexuallyisomorphic<br />

glomeruli in males and females, and pheromone activated<br />

the sexually-dimorphic macroglomerular complex in males. The NO<br />

response was concentration-dependent both <strong>for</strong> pheromone and <strong>for</strong><br />

linalool, a plant volatile. The NO-induced fluorescence was reduced by<br />

bath-applying NO signaling antagonists. These results show clearly<br />

that NO is produced in the AL in response to odorant stimulation of the<br />

antennae. The resulting NO may then stimulate soluble guanylyl<br />

cyclase in a subset of AL neurons to affect the processing of olfactory<br />

in<strong>for</strong>mation. Supported by NIH-NIDCD DC04292<br />

4 Slide [ ] Olfactory Bulb Physiology<br />

RESPONSES IN THE MOUSE MAIN AND ACCESSORY<br />

OLFACTORY BULBS TO GENERAL ODORANTS AND<br />

PHEROMONES REVEALED BY FMRI<br />

Xu F. 1, Liu N. 2, Kida I. 3, Schafer M. 4, Rothman D.L. 2, Hyder F. 3,<br />

Restrepo D. 4, Shepherd G.M. 1 1Neurobiology, Yale University, New<br />

Haven, CT; 2Yale University, New Haven, CT; 3Diagnostic Radiology,<br />

Yale University, New Haven, CT; 4University of Colorado Health<br />

<strong>Sciences</strong> Center, Denver, CO<br />

High resolution fMRI has been used to study the responses in the<br />

mouse main and accessory olfactory bulb (MOB and AOB) to general<br />

odors and pheromones. General odorants elicited strong signals across<br />

the MOB; a few odors were also able to activate the AOB. Urine, a rich<br />

source of pheromones, stimulated both the MOB and the AOB. The<br />

patterns in the MOB and AOB were distinctly different <strong>for</strong> urine and<br />

amyl acetate. The patterns of urines from different strains shared<br />

significant domains, and correlated well with other mapping methods.<br />

For both general odors and pheromones, the pattern intensity in the<br />

MOB increased with concentration, with the pattern topography<br />

remaining similar. In the AOB, most odors did not significantly<br />

activate any region at low concentration, but did at high concentration.<br />

Both pattern intensity and topography were concentration dependent in<br />

the AOB. However, pheromones stimulated the AOB at lower<br />

concentrations. Interestingly, the most activated regions were not<br />

significantly affected by concentration, making the topography of the<br />

pattern at different concentrations similar, as in the MOB. The<br />

response time course was slower in the AOB than in the MOB. The<br />

results give insights into the similarities and differences between<br />

responses of the MOB and AOB to general odors and pheromones.<br />

The research is supported by MURI and NIH grants DC-00086<br />

(GMS) and DC-03710(FH)


5 Slide [ ] Olfactory Bulb Physiology<br />

MAIN OLFACTORY BULB DETECTION OF SOCIAL<br />

RECOGNITION CUES IN MOUSE URINE<br />

Lin D. 1, Katz L.C. 1 1Neurobiology, Duke University, Durham, NC<br />

In rodents, chemicals present in urine play a critical role in<br />

reproductive behaviors and mediating aggressive interactions. Both<br />

behavioral tests and immediate early gene expression indicate that the<br />

main olfactory bulb (MOB) participates in individual identification.<br />

Little is known about how urine responsive neurons are organized in the<br />

MOB, or which semiochemicals in urine carry important in<strong>for</strong>mation.<br />

To address these issues, we examined extracellular activity of mitral<br />

cells in the mouse MOB in response to volatile components of urine<br />

from different sexes and strains of mice. Recordings from more than<br />

one thousand cells revealed that less than 5% of neurons responded<br />

selectively to different sources of urine. The urine responsive cells were<br />

confined to a small region in the ventral and lateral parts of the MOB.<br />

To ascribe a role to specific chemicals present in urine, we combined<br />

gas chromatography with extracellular recording in the MOB, so that<br />

the activity of individual mitral cells could be monitored while<br />

delivering the separated urinary components. The majority of neurons<br />

which showed an excitatory response to complete urine responded<br />

identically to just a single peak out of the hundreds present in the gas<br />

chromatogram of mouse urine. Surprisingly, over 20% of the urine<br />

responsive cells responded to one particular component, which has been<br />

implicated as a mouse pheromone. Preliminary behavioral data indicate<br />

the concentration of this pheromone in male mouse urine significantly<br />

correlates with the duration of female mice investigation of the urine.<br />

Thus, mitral cells selectively respond to individual components in<br />

mouse urine, and a disproportionate number of these neurons are<br />

involved in detecting a volatile pheromone.<br />

6 Slide [ ] Olfactory Bulb Physiology<br />

SPATIO-TEMPORAL FIRING RATE INTERACTIONS<br />

AMONGST AN ENSEMBLE OF MITRAL/TUFTED CELLS<br />

MAY SUBSERVE ODORANT DISCRIMINATIONS.<br />

Lehmkuhle M.J. 1, Normann R.A. 1, Maynard E.M. 1 1Bioengineering,<br />

University of Utah, Salt Lake City, UT<br />

Populations of output neurons in the mammalian olfactory bulb (OB)<br />

exhibit distinct spatial and temporal activation patterns when stimulated<br />

with enantiomers of the same odorant. If, as has recently been<br />

suggested, a two-stage network of inhibitory and excitatory centersurround<br />

connections link spatially distributed glomerular activity with<br />

underlying mitral and tufted neurons within olfactory bulb, then it is<br />

likely that these interactions affect the ensemble response to odorant<br />

stimulation. Here we investigate in the anesthetized rat the rate<br />

response temporal kinetics in neuronal ensembles by comparing pairwise<br />

response similarity of a population of single- and multi-unit<br />

mitral/tufted cells aligned to breathing in the presence and absence of<br />

enantiomers of limonene. Such comparisons can be used to quantify<br />

differences in the kinetics of the response of one neuron from other<br />

neurons in the recorded population. It is shown that aligning unit<br />

responses to the inhalation phase of the animals´ breathing tends to<br />

synchronize unit responses. Instances of high-similarity and<br />

dissimilarity within the population are enantiomer-specific with these<br />

instances occurring <strong>for</strong> pairs of units recorded over electrode<br />

separations of 400 µm – 1.7 mm. These results support the centersurround<br />

hypothesis and indicate that spatio-temporal firing rate<br />

interactions amongst an ensemble of OB neurons produce odorant<br />

representations beyond simple firing rates that may subserve odorant<br />

discrimination.<br />

2<br />

7 Slide [ ] Olfactory Bulb Physiology<br />

PRESYNAPTIC CENTER-SURROUND INHIBITION SHAPES<br />

ODORANT-ELICITED INPUT TO THE MOUSE OLFACTORY<br />

BULB<br />

Vucinic D. 1, Cohen L.B. 1, Kosmidis E. 1 1Cellular & Molecular<br />

Physiology, Yale University, New Haven, CT<br />

Synaptic transmission from olfactory primary afferents onto neurons<br />

of the olfactory bulb is modulated by presynaptic GABA receptors<br />

B<br />

(e.g. Wachowiak and Cohen, 1999; Ennis et al., 2001; Wachowiak et<br />

al., 2002). To study the role of this modulatory circuit in vivo we<br />

recorded the fluorescent signal from Calcium Green-1 dextran, loaded<br />

into the axon terminals of olfactory receptor neurons, in response to<br />

odorant presentations to the nose. We bath-applied agonists and<br />

antagonists of the GABA receptor onto the olfactory bulb and looked<br />

B<br />

<strong>for</strong> an effect of these agents on the amplitude, spatial map and time<br />

course of odorant-elicited signals. The GABA antagonist CGP46381<br />

B<br />

caused a significant and long-lasting increse in the average amplitude of<br />

glomerular activation in most preparations. Large changes in signal<br />

amplitude were accompanied by a change in the input map such that<br />

many weakly activated glomeruli underwent relatively larger increases<br />

in signal amplitude than the strongy activated ones. We find that the<br />

magnitude of this effect correlates with how strongly activated the<br />

surround of a glomerulus is. This finding indicates that a <strong>for</strong>m of<br />

GABA-mediated center-surround inhibition modifies the sensory input<br />

map even be<strong>for</strong>e the first synaptic transmission, increasing its contrast.<br />

Supported by NIH grants NS07455 and DC05259.<br />

8 Slide [ ] Olfactory Bulb Physiology<br />

LONG-TERM ODOR EXPOSURE INCREASES SURVIVAL<br />

AND FUNCTIONAL INTEGRATION OF INTERNEURONS IN<br />

THE OLFACTORY BULB.<br />

Mirich J. 1, Illig K.R. 1, Brunjes P.C. 1 1Psychology, University of<br />

Virginia, Charlottesville, VA<br />

While our previous work indicates that activity is critical <strong>for</strong> the<br />

survival of bulbar interneurons, the rules by which new neurons become<br />

integrated into existing neural circuits remain unclear. If olfactory<br />

activity guides the functional integration and survival of newly born<br />

cells, then repeated exposure to an odorant should increase the number<br />

of cells that participate in the coding of that odor. Adult subjects<br />

received daily injections of BrdU <strong>for</strong> 5 days to label a cohort of newly<br />

born neurons. They were then exposed daily to menthyl isovalerate, ßpinene,<br />

butyric acid, or mineral oil (control) <strong>for</strong> 3 weeks. For half of the<br />

animals in each group, the odorant was paired with Froot Loops to<br />

increase salience. On the last day, animals were exposed to ultra-pure<br />

odorants intermittently <strong>for</strong> 1hr, and tissue stained <strong>for</strong> BrdU and Fos<br />

protein. Long-term odor exposure doubled the survival of interneurons<br />

compared with controls. Double-label immunohistochemistry showed a<br />

statistically significant increase (44%; p < 0.05) in BrdU/Fos positive<br />

cells, demonstrating integration of newly born cells during the training<br />

period. Additional analyses confirmed that new neurons integrated<br />

specifically into the areas of the bulb activated by the odorant<br />

presented. This study demonstrates that survival and integration of<br />

newly born bulbar neurons depends on the odor environment present<br />

during their development. We conclude that repeated activation of<br />

synaptic ensembles might change neural requirements, thereby<br />

influencing the cellular milieu. Supported by NIH grants DC0338<br />

(NIDCD) and HD07323 (NICHD).


9 Slide [ ] Olfactory Bulb Physiology<br />

A LIMK DISEASE MODEL REVEALS DEFECTS IN<br />

GLOMERULAR DEVELOPMENT AND OLFACTORY<br />

NEURODEGENERATION<br />

Hing H.K. 1, Ang L. 2, Yao Y. 2, Uemura T. 3, Keshishian H. 4 1Cell and<br />

Structural Biology, University of Illinois at Urbana-Champaign,<br />

Urbana, Illinois; 2Cell and Structural Biology, University of Illinois at<br />

Urbana-Champaign, Urbana, IL; 3Kyoto University Hospital, kyoto, IL;<br />

4Molecular Cell and Developmental Biology, Yale University, new<br />

haven, CT<br />

The molecular mechanism by which the precise connectivity of the<br />

olfactory map is <strong>for</strong>med is poorly understood. We have investigated the<br />

role of the Lim Kinase (Limk) gene in the Drosophila olfactory map<br />

development. Limk has also been at the center of much attention lately<br />

because its loss is implicated in the human mental retardation disease,<br />

Williams Syndrome. We have created a Limk disease model by<br />

knocking out and overexpressing the fly Limk gene. At the<br />

neuromuscular junction, loss of Limk leads to enlarged synapses while<br />

expression of a hyperactive Limk mutant (LimkKd ) leads to stunted<br />

synapse. Expression of LimkKd in the olfactory receptor neurons<br />

(ORNs) results in the <strong>for</strong>mation of numerous ectopic glomeruli. The<br />

ability of the LimkKd to induce ectopic glomeruli is abolished by either<br />

mutations in the Pak gene or overexpression of cofilin gene. This result<br />

indicates that Pak, Limk and cofilin function in a signaling pathway to<br />

regulate synaptogenesis in the fly antennal lobes. Interestingly, loss of<br />

Limk also leads to an adult-onset, progressive loss of ORN axons from<br />

the antennal lobes. This adult-onset neurodegeneration phenotype is<br />

mimicked by increased cofilin expression. In summary, Limk per<strong>for</strong>ms<br />

two functions in vivo, a developmental function in which it regulate<br />

synapse development and an adult function in which it protects neurons<br />

from premature demise by keeping cofilin activity in check.<br />

10 Symposium [ ] Developmental Regulatory Genes in the<br />

Taste and Olfactory Systems<br />

MAKING A NEURON: PRONEURAL BHLH FACTORS<br />

DURING RETINAL AND OLFACTORY DEVELOPMENT<br />

Vetter M.L. 1 1Neurobiology and Anatomy, University of Utah, Salt<br />

Lake City, UT<br />

Basic helix-loop-helix (bHLH) transcription factors play a central<br />

role in regulating the acquisition of neuronal cell fate in sensory tissues<br />

such as the olfactory epithelium and neural retina. For example, we<br />

identified the bHLH factor Xath5 in Xenopus laevis and demonstrated<br />

that it can regulate neuronal differentiation during retina and olfactory<br />

placode development. In general, proneural bHLH factors are both<br />

necessary and sufficient to promote the neural fate and do so by<br />

activating a core program of neuronal differentiation in progenitors. We<br />

have per<strong>for</strong>med a differential screen to define the genes that are directly<br />

regulated by proneural bHLH factors in Xenopus and are examining<br />

how these genes contribute to the process of neuronal differentiation.<br />

In addition, specific factors act to regulate the onset of proneural<br />

gene expression during development. Frizzled 5, which is a<br />

transmembrane receptor that binds to wnt ligands, is exclusively<br />

expressed in the developing neural retina during early eye development<br />

in Xenopus laevis. We used antisense morpholino oligonucleotides to<br />

disrupt Frizzled 5 expression and found that this receptor regulates the<br />

onset of neurogenesis during retinal development. In the absence of<br />

normal Frizzled 5 expression, retinal progenitors fail to activate the<br />

expression of genes required <strong>for</strong> neural competence and subsequently<br />

fail to initiate proneural gene expression on schedule. Thus we can<br />

place proneural bHLH factors in a genetic cascade that leads to the final<br />

acquisition of a neural fate.<br />

3<br />

11 Symposium [ ] Developmental Regulatory Genes in the<br />

Taste and Olfactory Systems<br />

COMMON MOLECULAR SIGNALS REGULATING<br />

PROGRESSION THROUGH THE NEURONAL LINEAGE IN<br />

OLFACTORY AND OTHER SENSORY EPITHELIA<br />

Calof A.L. 1, Beites C. 1, Crocker C. 1, Hayashi H. 1, Kim J. 1, Silman E. 1,<br />

Santos R. 1, Kawauchi S. 1 1Anatomy & Neurobiology, University of<br />

Cali<strong>for</strong>nia, Irvine, Irvine, CA<br />

To understand how signaling molecules regulate the generation of<br />

neurons from proliferating stem cells and neuronal progenitors in the<br />

mammalian nervous system, we have focused on studies of<br />

neurogenesis in the olfactory epithelium (OE) of the mouse. By<br />

analyzing neurogenesis in the OE of normal animals and mouse<br />

developmental mutants, we have identified distinct stages of stem and<br />

transit amplifying progenitor cells in the differentiation pathway of<br />

olfactory receptor neurons (ORNs); each of these cell stages can be<br />

identified by distinct molecular marker(s). Studies of progenitor cells in<br />

developing and regenerating OE have led to the understanding that (1)<br />

each progenitor cell type is regulated by signals produced both within<br />

the OE itself and by its underlying stroma; (2) the same regulatory<br />

gene(s) play critical roles in neurogenesis in different regions of the<br />

primary olfactory pathway; (3) the same signaling molecules that<br />

regulate neurogenesis in OE likely play similar roles in other sensory<br />

epithelia. Supported by grants to ALC from the NIH (DC03583 and<br />

HD38761) and March of Dimes.<br />

12 Symposium [ ] Developmental Regulatory Genes in the<br />

Taste and Olfactory Systems<br />

SHH SIGNALING AND TASTE BUD MAINTENANCE IN THE<br />

ADULT MOUSE<br />

Miura H. 1, Kusakabe Y. 1, Tetsuya O. 1, Ninomiya Y. 2, Hino A. 1<br />

1National Food Research Institute, Tsukuba, Japan; 2Kyushu<br />

University, Fukuoka, Japan<br />

Continuous cell renewal in adult mammalian taste buds implies the<br />

requirement of inductive signals <strong>for</strong> growth and differentiation of taste<br />

bud precursor cells. And they should be dependent on the taste nerve.<br />

We have previously reported the expression of Sonic hedgehog (Shh) in<br />

the basal cells of taste buds. Various inductive events are mediated by<br />

the Shh signal in animal development. In adult taste epithelium, the<br />

expression of Patched1 (Ptc), Shh receptor, was observed in the<br />

surrounding region of the Shh expression, where mitotic cells<br />

contributing to taste buds may be distributed. Moreover, Nkx2.2<br />

expression was found in taste buds: Nkx2.2 is transiently expressed in<br />

the neuronal precursor cells in the ventral region of neural tube, being<br />

activated by Shh signal. The Nkx2.2-expressing cells in the taste buds<br />

were found to express Mash1, a marker <strong>for</strong> the neuronal precursor cells.<br />

These observations suggest that Shh signaling may be involved in the<br />

taste bud maintenance in the adult mouse. In addition, the denervation<br />

revealed the strong nerve dependency of the basal cell-specific Shh<br />

expression in the taste bud. A quick loss of Shh after denervation led<br />

the decrease of Ptc expression. The decrease might be involved in the<br />

arrest of precursor cell proliferation after denervation, which causes the<br />

taste bud disappearance. Shh signaling and the regulatory gene<br />

expressions in taste buds will be discussed in relation to the taste cellspecific<br />

genes. This work was supported by Bio-Oriented Technology<br />

Research Advancement Institution of Japan.


13 Poster [ ] Taste Hedonics & Psychophysics<br />

UNDERSTANDING VEGETABLE ACCEPTANCE: ROLE OF<br />

EARLY EXPERIENCE<br />

Kennedy J.M. 1, Mennella J.A. 1, Beauchamp G.K. 1 1Monell Chemical<br />

Senses Center, Philadelphia, PA<br />

Although infants differ substantially in their acceptance of foods<br />

during weaning, the source of such differences remains a mystery.<br />

Recently, we have identified a particularly apt system to explore these<br />

issues: the inherent flavor variations characteristic of infant <strong>for</strong>mulas.<br />

Within each of these categories of <strong>for</strong>mulas are a number of varieties<br />

that differ among themselves in <strong>for</strong>mulation and flavor but the<br />

differences between the categories, and, in particular, between the<br />

hydrolysate and milk-based varieties in sensory quality (flavor) are<br />

striking and profound. The present study tested the hypothesis that the<br />

type of <strong>for</strong>mula fed to infants would influence their acceptance of<br />

vegetables that shared a similar flavor note (e.g., sulfur volatiles) with<br />

the <strong>for</strong>mula (e.g., hydrolysate <strong>for</strong>mula). In counterbalanced order, we<br />

evaluated 87 infants´ acceptance of pureed carrot on one testing day and<br />

pureed broccoli on the other. Infants who were feeding a hydrolysate<br />

<strong>for</strong>mula consumed significantly less broccoli relative to carrots when<br />

compared to those who were currently fed milk based <strong>for</strong>mulas. Such<br />

findings are consistent with previous research that demonstrated a<br />

sensory specific satiety following repeated exposure to a particular<br />

flavor in either <strong>for</strong>mula or mothers´ milk in the short term. This<br />

research was supported by NIH Grant HD37119.<br />

14 Poster [ ] Taste Hedonics & Psychophysics<br />

ANALGESIC EFFECTS OF INTRAORAL SUCROSE: THE<br />

MORE THEY LIKE SWEET TASTE, THE BETTER IT<br />

WORKS?<br />

Pepino M.Y. 1, Kennedy J.M. 1, Mennella J.A. 1 1Monell Chemical Senses<br />

Center, Philadelphia, PA<br />

During infancy and childhood, preference <strong>for</strong> sweet tastes is<br />

heightened and sweet-tasting substances can be analgesics. The goal of<br />

the study was to evaluate individual differences in sweet preferences<br />

and to determine whether such differences are related to sucrose´s<br />

analgesic effects in 5- to 10-year-old children and their mothers. To this<br />

aim, the preferred level of sucrose was determined by using a <strong>for</strong>cedchoice,<br />

paired comparison, tracking procedure, and the analgesic effect<br />

of sweet taste was determined by the Cold Pressor Test. As a group,<br />

children preferred significantly higher concentrations of sucrose than<br />

mothers. Ethnic differences in sweet preferences were observed in both<br />

children and adults such that Blacks preferred significantly higher<br />

concentrations when compared to Whites. Regardless of race, children<br />

who preferred high sweet concentrations kept their hand in the cold<br />

water significantly longer when sucrose was held in their mouths when<br />

compared to water (p=0.002). These findings suggest that the analgesic<br />

effects of sweet tastes may be more pronounced in those pre-pubertal<br />

children who have heightened sweet preference. This research was<br />

supported by NIH Grants AA09523 and HD37119.<br />

4<br />

15 Poster [ ] Taste Hedonics & Psychophysics<br />

INFLUENCE OF CONCENTRATION ON TASTE-TASTE<br />

INTERACTIONS IN FOODS BY ELDERLY AND YOUNG<br />

Mojet J. 1, Heidema J. 2, Christ-Hazelhof E. 2 1Consumer & Market<br />

Insight, Agrotechnology and Food Innovations, Wageningen,<br />

Netherlands; 2Unilever Research and Development, Vlaardingen,<br />

Netherlands<br />

An increase in concentration of one of the tastants in a `real food´<br />

might not only effect the perception of the taste quality of the<br />

manipulated tastant, but also the other perceivable taste qualities. The<br />

influence of concentration increase of sodium chloride, potassium<br />

chloride, sucrose, aspartame, acetic acid, citric acid, caffeine, quinine<br />

HCl, monosodium glutamate (MSG), and inosine 5´-monophosphate<br />

(IMP) on the other perceivable taste qualities was studied in different<br />

foods.Twenty-one young (19-33 years) and 21 older subjects (60-75<br />

years) rated the saltiness, sweetness, sourness, bitterness and umami<br />

taste of the food stimuli on 9-point scales. Repeated measures and<br />

multivariate analysis showed that an increasing concentration of sodium<br />

and potassium chloride diminished the sweetness more <strong>for</strong> the young<br />

than <strong>for</strong> the elderly, but enlarged the bitterness, sourness and umami<br />

taste in tomato soup more <strong>for</strong> the elderly than <strong>for</strong> the young. The<br />

saltiness of ice tea was decreased with an increase in sucrose, while a<br />

larger decrease in bitterness was found <strong>for</strong> the young than <strong>for</strong> the<br />

elderly. An increase in sucrose or aspartame concentration in ice tea<br />

also decreased the sourness. No influence was shown by an increment<br />

of acetic acid or citric acid in mayonnaise. The increasing concentration<br />

of caffeine and quinine induced a decrease in sweetness of the<br />

chocolate drink. The increase in MSG showed an increase in the<br />

saltiness of broth, whereas an increase in IMP led to a decrease in<br />

saltiness and an increase in sweetness. Young subjects took advantage<br />

of their sense of smell in the no noseclip condition.<br />

16 Slide [ ] Taste Hedonics & Psychophysics<br />

ACCOUNTING FOR BETWEEN-SUBJECT VARIANCE IN<br />

DISCRIMINATION AND PREFERENCE TASKS<br />

Delwiche J. 1, Liggett R. 1 1Food Science and Technology, Ohio State<br />

University, Columbus, OH<br />

The binomial statistic is typically used to determine the number of<br />

correct discriminations needed to indicate a significant difference<br />

between items. This statistic does not account <strong>for</strong> differences between<br />

subjects, making it inappropriate to combine responses across subjects<br />

and repetitions, which is often something researchers would like to do.<br />

The beta-binomial model is able to account <strong>for</strong> between-subject<br />

variance (measured by gamma), making such analyses possible. This<br />

study examined panel overdispersion (gamma) <strong>for</strong> one group of<br />

subjects (53-58 subjects per group, depending on stimuli set)<br />

per<strong>for</strong>ming two replications of both paired comparisons (2AFC) and<br />

paired preferences on the same stimuli set. Stimuli tested included fruitflavored<br />

beverages (with different sucrose levels), and snack foods<br />

(with different fat contents). Results showed that significant<br />

overdispersion with one task were not predictive of overdispersion in<br />

the other. Further, the stability of gamma across discrimination methods<br />

(2AFC, 3AFC, triangle, and duo-trio) <strong>for</strong> a group of 103 subjects was<br />

examined. Stimuli were cherry-flavored fruit beverages at two different<br />

sucrose levels, and order of the discrimination tasks was counterbalanced<br />

across subjects. Results indicated that gamma was largely<br />

consistent across the 2AFC, 3AFC and triangle tasks, but it was higher<br />

in the duo-trio task. In all cases, the use of the beta-binomial model<br />

allowed <strong>for</strong> the combining of discriminations across subjects and<br />

replications, increasing the discrimination power achieved <strong>for</strong> a given<br />

panel size. This project was self-funded.


17 Poster [ ] Taste Hedonics & Psychophysics<br />

RESPONSES OF PROP TASTER GROUPS TO VARIATIONS IN<br />

TASTES AND ORAL IRRITATION WITHIN A BEVERAGE<br />

Prescott J. 1, Campbell H. 2, Roberts C. 2 1Psychology, James Cook<br />

University, Cairns, QLD, Australia; 2Sensory Science Research Centre,<br />

University of Otago, Dunedin, New Zealand<br />

Despite considerable evidence that variations in sensitivity to the<br />

bitterness of 6-n-propylthiouracil (PROP) are also reflected in<br />

responses to both other tastes and also chemesthetic qualities in<br />

solution, there has been little research examining the impact of PROP<br />

sensitivity on response to these sensory modalities in foods or<br />

beverages. The present study examined responses of PROP taster<br />

groups to systematic variations in sourness and oral irritation in a<br />

carbonated beverage. Taster groups were defined according to ratings<br />

of a 0.0032 M PROP solution using the labelled magnitude scale<br />

(LMS). Using the LMS, Ss rated the sweetness, sourness and oral<br />

irritation of carbonated fruit drinks that systematically varied in citric<br />

acid (0.32, 0.64, 1.28, 2.56% w/v) and CO2 (25, 50, 75 psi)<br />

concentrations. Ratings of sourness as a function of citric acid, and<br />

irritation as a function of both citric acid and CO2 levels, were<br />

significantly different between groups, with highest ratings <strong>for</strong> STs and<br />

lowest <strong>for</strong> NTs. There were no group differences <strong>for</strong> sweetness ratings.<br />

These data are some of the first to show PROP taster group differences<br />

in tastes and irritation within a real product, and provide a basis <strong>for</strong><br />

reported differences of PROP groups in their hedonic responses to<br />

foods.<br />

18 Poster [ ] Taste Hedonics & Psychophysics<br />

GENETIC SENSITIVITY TO 6-N-PROPYLTHIOURACIL<br />

(PROP), AND PERCEPTION OF HIGH-INTENSITY<br />

SWEETENERS IN MODEL SOFT DRINKS<br />

Tepper B.J. 1, Zhao L. 1 1Food Science, Rutgers University, New<br />

Brunswick, NJ<br />

Intense sweeteners (e.g., saccharin, aspartame, and acesulfame-K)<br />

impart sweetness, bitterness and other aftertastes to foods. Individual<br />

variation in sensitivity to sweeteners is well known. Those who are<br />

genetically sensitive to the bitterness of PROP perceive more sweetness<br />

and bitterness from saccharin. Findings <strong>for</strong> other sweeteners are less<br />

clear. This study examined the role of PROP taster status in perception<br />

and acceptance of intense sweeteners in citrus-flavored model soft<br />

drinks. Young adults were classified as non-tasters (NT; n=29) or<br />

supertasters (ST; n=30) of PROP using a filter paper screening method<br />

(Zhao et al., 2003). The following sweeteners were used: 10% and 8%<br />

high fructose corn syrup (HFCS) (controls), sucralose (SUC),<br />

aspartame (ASP), acesulfame-K (ACE), ASP/ACE and SUC/ACE.<br />

Subjects rated the intensity of nine attributes with a 15-cm line scale<br />

and liking with the Labeled Affective Scale (LAM). ST perceived more<br />

bitterness from SUC/ACE and 8% HFCS (p=0.05); ACE was<br />

marginally more bitter to ST (p=0.06). ST also perceived more<br />

persistence of sweetness across all sweeteners (p=0.05). 3-dimensional<br />

models best described the perceptions of the sweeteners. For ST the<br />

dimensions were bitterness (43% variance), persistence of sweetness<br />

(22%), and carbonation (11%). For NT the dimensions were: sweetness<br />

and citrus flavor (37%), bitterness (24%), and carbonation/thickness<br />

(11%). Liking was uniquely related to low bitterness <strong>for</strong> NT but was<br />

related to multiple attributes <strong>for</strong> ST. These data suggest that ST<br />

experience intense sweeteners differently than NT but these differences<br />

play a modest role in soft drink acceptance. Supported by McNeil<br />

Nutritionals.<br />

5<br />

19 Poster [ ] Taste Hedonics & Psychophysics<br />

6-N-PROPYLTHIOURACIL (PROP) BITTERNESS AND<br />

TASTES FROM ALCOHOLIC AND NON-ALCOHOLIC<br />

BEVERAGES IN OF-AGE UNDERGRADUATES<br />

Lanier S. 1, Hayes J. 2, Duffy V.B. 1 1Dietetics, University of Connecticut,<br />

Storrs, CT; 2Nutritional Science, University of Connecticut, Storrs, CT<br />

Previous research has shown associations between taste genetics and<br />

alcohol behaviors. We tested the taste genetic and alcohol intake<br />

relationship and linked this relationship to oral sensations from<br />

alcoholic beverages. Taste genetics was measured via perceived PROP<br />

bitterness; nontasters taste PROP as mildly bitter while supertasters as<br />

intensely bitter. Sixty-two adults (mean age=22 years) used the general<br />

Labeled Magnitude Scale to rate PROP bitterness and oral sensations<br />

from and preference <strong>for</strong> sampled Pilsner beer, scotch, strong black<br />

coffee and unsweetened grapefruit juice. Subjects also completed<br />

validated questionnaires on behaviors toward alcohol. Data were<br />

analyzed with regression analyses. Those who tasted greater PROP<br />

bitterness reported all beverages as more bitter and less preferred except<br />

beer. Neither PROP bitterness nor beer bitterness were strong<br />

predictors of beer preference. PROP nontasters tasted scotch as less<br />

bitter but more sweet than did supertasters. Those who tasted scotch as<br />

more bitter consumed less alcohol; PROP bitterness only tended to<br />

associate with alcohol intake. Beer sensations did not associate<br />

significantly with alcohol intake. Summary: Nontasters get less<br />

negative (bitter) and more positive (sweet) tastes from alcohol than<br />

supertasters. Sensory differences translated into greater liking <strong>for</strong> bitter<br />

beverages by nontasters, except <strong>for</strong> beer. Positive social influences to<br />

drink beer on college campuses may override negative oral sensations<br />

to hinder intake. (NRICGP/USDA 2002-00788 funded)<br />

20 Poster [ ] Taste Hedonics & Psychophysics<br />

PERSONALITY TRAITS AND PROP SENSITIVITY<br />

White T.L. 1, Longo M. 2 1Psychology, Le Moyne College, Syracuse, NY;<br />

2Education, Syracuse University, Syracuse, NY<br />

Although theories of personality vary widely, most trait theories<br />

include the dimension of Introversion-Extroversion. Physiological<br />

differences are thought to exist between people who are represented by<br />

the extremes on this dimension, with introverts experiencing a high<br />

level of baseline arousal. This higher arousal level suggests that<br />

introverts experience elevated sensitivity to perceptual stimuli,<br />

including sour tastes (Eysenck, 1976), as compared to extroverts. As the<br />

ability to taste PROP can also affect taste sensitivity, the current study<br />

investigated the relationship between PROP sensitivity and personality<br />

traits, particularly introversion. The NEO-PI, a personality measure that<br />

evaluates 5 personality traits, was administered to 50 college students<br />

(39 women, 11 men). After completing the personality test, participants<br />

were given a PROP sample and asked to rate its intensity on a labeled<br />

magnitude scale (LMS; Green, 1993). No correlation was found<br />

between introversion, conscientiousness, or openness to experience and<br />

PROP tasting ability. However, a relationship was seen between the<br />

domain trait of neuroticism and intensity ratings of PROP on the LMS<br />

(r=0.346, p=0.014), in particular those facets concerned with anxiety<br />

and depression. Although agreeableness was not overall associated, the<br />

facet concerned with trust was inversely (r=-0.336, p=0.017) associated<br />

with LMS Scores. These associations suggest that personality traits may<br />

interact with genetic abilities as partial explanation <strong>for</strong> individual<br />

differences food preferences.


21 Poster [ ] Taste Hedonics & Psychophysics<br />

OROSENSORY AND GENETIC TASTE (GT) MARKERS<br />

PREDICT ALCOHOL INTAKE ACROSS AGE COHORTS<br />

Hayes J.E. 1, Chapo A. 2, Bartoshuk L. 2, Duffy V.B. 3 1Nutritional<br />

<strong>Sciences</strong>, University of Connecticut, Storrs, CT; 2Surgery, Yale<br />

University, New Haven, CT; 3Dietetics, University of Connecticut,<br />

Storrs, CT<br />

Supertasters, identified by heightened 6-n-propylthiouracil (PROP)<br />

bitterness and fungi<strong>for</strong>m papilla (FP) number, report greater negative<br />

and less positive sensations from alcohol, possibly hindering intake.<br />

Work by others (eg, Kranzler et al) show positive associations between<br />

oral sensations and intake, findings that contradict GT-mediated effects.<br />

To characterize age, sensory, and GT effects on intake, 99 healthy<br />

women (aged 20-84) used the general Labeled Magnitude Scale to rate<br />

PROP bitterness, regional and whole mouth intensity of prototypical<br />

tastants and ethanol on the tongue tip. Intake was assessed via a<br />

frequency interview and FP by videomicroscopy. In young subjects,<br />

lower PROP bitterness and greater whole mouth taste were significant<br />

predictors of greater alcohol intake, yet a spatial taste measure<br />

diminished the predictive ability of whole mouth taste. Across the entire<br />

sample, age (young>aged), FP number (higher number, less intake) and<br />

whole-mouth taste responses (greater intensity, more intake) were<br />

significant predictors of intake. Cross quality whole-mouth and chorda<br />

tympani, but not circumvallate, intensity showed age-related loss. In<br />

summary, GT markers and additional taste markers contribute to<br />

predicting alcohol intake across age cohorts. Reasons <strong>for</strong> higher taste<br />

response associating with greater alcohol intake remain unclear. Spatial<br />

interactions among sensory nerves could heighten whole mouth<br />

orosensation (eg, sweetness) and change the sensory appeal of alcohol.<br />

(NRICGP/USDA 2002-00788)<br />

22 Poster [ ] Taste Hedonics & Psychophysics<br />

INFLUENCE OF FATTY ACID SENSITIVITY ON TASTE<br />

PERCEPTION<br />

Smeets M. 1, Weenen H. 2, De Wijk R. 2, Mojet J. 1, Westerterp M. 3<br />

1Agrotechnology & Food Innovations, Wageningen, Netherlands;<br />

2Wageningen Centre <strong>for</strong> Food <strong>Sciences</strong>, Wageningen, Netherlands;<br />

3Human Biology, Maastricht University, Maastricht, Netherlands<br />

Following previous work (Gilbertson et al., Kamphuis et al.)<br />

demonstrating chemosensitivity to fatty acids in the taste system, we<br />

investigated whether sensitivity to fatty acids in humans:<br />

1. is influenced by olfaction<br />

2. is associated with differences in taste perception in general.<br />

Methods: A group of (linoleic acid) LA-tasters (≥ 9 out of 10 correct<br />

detections of 10 µM LA) was compared to a non-taster group (n=20<br />

total). To investigate 1) triangle tests were conducted using 0 % fat<br />

mayonaise with and without LA (10 µM), with nose clips on and off.<br />

To investigate 2) taste profiles were generated <strong>for</strong> various<br />

products/solutions containing e.g. LA and basic tastants, with and<br />

without noseclips.<br />

Results:<br />

1. LA-Tasters showed above-chance per<strong>for</strong>mance without noseclips<br />

(p=0.04), but not with noseclips. Non-tasters did not per<strong>for</strong>m above<br />

chance.<br />

2. There was a main effect of taster status (p


25 Poster [ ] Taste Hedonics & Psychophysics<br />

INHIBITION OF BITTER TASTE BY ZINC AND NA-<br />

CYCLAMATE<br />

Keast R.S. 1, Breslin P.A. 2 1Monell Chemical Senses Center,<br />

Philadelphia, PA; 2Psychophysics, Monell Chemical Senses Center,<br />

Philadelphia, PA<br />

Previous research has shown that zinc is a potent inhibitor of the<br />

bitterness of quinine-HCl (Keast, 2003), and the sweet taste of most<br />

sweeteners (except Na-cyclamate) (Keast et al., 2004). We investigated<br />

the influence of zinc on perceived bitterness of six structurally<br />

divergent bitter compounds. We also determined the impact of a<br />

combination of zinc & Na-cyclamate or zinc & sucrose on the bitterness<br />

of specific compounds.<br />

In experiment one, six diverse bitter tasting compounds (Quinine-<br />

HCl (QHCl), Tetralone (TET), Sucrose octaacetate, Dextromethorphan,<br />

Denatonium Benzoate (DB), and Pseudoephedrine (PSE)) were<br />

matched <strong>for</strong> `moderate´ bitterness intensity. To these compounds one<br />

of five salts was added: 25 or 300mM Na acetate, 25mM Mg sulfate,<br />

25mM Mg acetate, and 25mM Zn sulfate. All possible binary<br />

combinations of bitter compounds and salts were tested. In experiment<br />

two, the influence of Zn sulfate on bitter-sweet mixtures was<br />

investigated. The bitter compounds used were DB and PSE, and the<br />

sweet compounds were sucrose and Na-cyclamate.<br />

Zn sulfate, 300mM Na acetate, and Mg acetate significantly<br />

inhibited bitterness (p


29 Poster [ ] Taste: Fats<br />

THE GUSTATORY SENSATION FROM FREE FATTY ACID<br />

Kawai T. 1, Nishiduka T. 1, Kajii Y. 2, Shingai T. 2, Kuwasako T. 3, Hirano<br />

K. 3, Yamashita S. 3, Kawada T. 1, Fushiki T. 1 1Graduate School of<br />

Agriculture, Kyoto University, Kyoto, Kyoto, Japan; 2Graduate School<br />

of Medical and Dental <strong>Sciences</strong>, Niigata University, Niigata, Niigata,<br />

Japan; 3Graduate School of Medicine, Osaka University, Suita, Osaka,<br />

Japan<br />

Fat in food often increases the palatability of food, even though it<br />

does not give us obvious taste by itself. Some researches have reported<br />

that taste cells recognize free fatty acid, a hydrolysate from common<br />

dietary fat, though the free fatty acid has not be defined as one of<br />

tastants. Its perceptual mechanism is still unclear. We demonstrated<br />

immunohistochemically that CD36, one of fatty acid transport proteins,<br />

was localized on the apical part of the taste cells isolated from rat<br />

vallate papillae. This localization supports the hypothesis that CD36<br />

plays a role in the oral recognition of dietary fat. In this study, we used<br />

CD36-null mice <strong>for</strong> behavioral studies and nerve recordings and<br />

compared with the wild type mice that had the same background. The<br />

wild type mice preferred oleic acid solution to mineral oil solution in<br />

short-time two-bottle preference tests, but the CD36-null mice did not<br />

show the preference <strong>for</strong> either solution. Neural recording revealed that<br />

small but significant response to oleic acid was observed on the lingual<br />

branch of glossopharyngeal nerve in the wild type mice, but not in the<br />

CD36-null mice. These results suggest that gustatory signal of fat was<br />

conveyed via the lingual branch of glossopharyngeal nerve and that the<br />

CD36 molecule, which exists on the tongue innervated by the lingual<br />

branch of glossopharyngeal nerve, plays an important role in enjoying a<br />

taste of fatty acid.<br />

30 Poster [ ] Taste: Fats<br />

PROP TASTER STATUS AND PERCEPTION OF FATS AND<br />

FREE FATTY ACIDS<br />

Armstrong C.L. 1, Mattes R. 1 1Department of Foods and Nutrition,<br />

Purdue University, West Lafayette, IN<br />

PROP tasters, especially supertasters, are reportedly better able to<br />

perceive the level of fat in foods than PROP non-tasters. This study<br />

tested perception of fat in both foods and free fatty acids in water.<br />

PROP taster status was determined by the 5-solution method. To date,<br />

7 non-tasters, 12 tasters and 7 supertasters have been tested. Subjects<br />

were randomly presented 5 ml samples of milk, (0.00, 3.25, 10.00,<br />

18.00, 36.00% fat/wt), 5 ml of vanilla pudding (4.1, 8.2, 16.3, 24.4,<br />

29.7% fat/wt) and 5 gm of brownies (8.0, 16.0, 24.0, 31.8, 35.6%<br />

fat/wt) and rated the level of fat in duplicate using the gLMS. Milk and<br />

vanilla pudding were served at 7° C and brownies at 22° C. To mask<br />

texture, oleic and linoleic acids (0.40, 0.71, 1.267, 2.25, 4.00% wt/wt)<br />

were suspended in a solidified unflavored gelatin starch-thickened<br />

water mixture and 5 gm samples were served at 22° C. Stearic acid (0.5,<br />

1.0, 2.0, 4.0, 8.0% wt/wt) was suspended in starch-thickened water and<br />

5 ml samples were served at 71° C. Samples were randomly presented<br />

in duplicate and the level of fat rated using the gLMS. Subjects<br />

correctly ranked the fat levels (p < 0.0001) in all foods and free fatty<br />

acids solutions, but no differences were observed between PROP taster<br />

groups. These data demonstrate a taste component <strong>for</strong> free fatty acids,<br />

but no differences in fat perception based on PROP taster status. This<br />

research was funded in part by the Rose Marie Pangborn Scholarship to<br />

CLA and NIH Grant #DK045294.<br />

8<br />

31 Poster [ ] Taste: Fats<br />

FAT TASTE - ARE FREE FATTY ACIDS OR CONJUGATED<br />

DIENES THE EFECTIVE STIMULUS?<br />

Chalé A. 1, Burgess J.R. 2, Mattes R.D. 1 1Foods and Nutrition, Purdue<br />

University, W Lafayette, IN; 2Foods & Nutrition, Purdue University, W<br />

Lafayette, IN<br />

"Fattiness" is hypothesized to be a basic taste quality, most<br />

efficiently elicited by long-chain poly- and monounsaturated fatty acids.<br />

Free fatty acids (FFA) have been prepared in solutions containing an<br />

emulsifier or thickening agent to mask textural cues or in foods.<br />

Whether the effective stimulus was the FFA or oxidation product has<br />

not been established. This work examined the effects of addition of<br />

EDTA and sonification (to reduce oxidation) on oxidation product<br />

concentration. Linoleic acid was prepared at a concentration of<br />

10mg/dl and mixed in a solution of 5% Acacia and de-mineralized<br />

water. The concentration of EDTA was 0.01%. Samples were sonified<br />

<strong>for</strong> 40 minutes. This preparation increased linoleic acid recovery 100X<br />

over a simple oil-water mixture. These results suggest extensive<br />

degradation of FFAs in previously used fat taste samples resulting in<br />

conjugated dienes that may be effective taste stimuli. This is being<br />

explored through psychophysical studies.<br />

32 Poster [ ] Vomeronasal Organ<br />

REGULATOR OF G-PROTEIN SIGNALING PROTEINS IN<br />

THE VOMERONASAL ORGAN OF GARTER SNAKES<br />

Wang D. 1, Liu W. 2, Chen P. 1, Halpern M. 3 1Biochemistry, Downstate<br />

Medical Center, Brooklyn, NY; 2Anatomy and Cell Biology, Downstate<br />

Medical Center, Brooklyn, NY; 3Anatomy & Cell Biology, Downstate<br />

Medical Center, Brooklyn, NY<br />

The response to prey chemoattractants in garter snakes is mediated<br />

by the vomeronasal (VN) system.The chemosignal transduction<br />

pathway in the VN epithelium involves the binding of agonist to its Gprotein<br />

coupled receptors leading to activation of Gi/o-proteins which<br />

in turn activate PLC, converting PIP2 to IP3 and DAG, and resulting in<br />

a transient cytosolic Ca2+ increase and changes in membrane potential.<br />

However, the desensitization mechanism so far remains unknown. We<br />

found that the chemoattractant-induced signal is modulated by two<br />

membrane-bound proteins, p42/44, which modulate the exchange of<br />

GTP <strong>for</strong> GDP on the Ga subunit. During recent years, regulators of Gprotein<br />

signaling (RGS proteins) have been identified. These proteins<br />

enhance the activity of GTPase intrinsic to G alpha subunits and have<br />

been shown, in a wide number of biological systems, to play an<br />

important role in modulating agonist-induced signals. Using<br />

commercially available RGS antibodies, we detected the expression of<br />

several isotypes of RGS proteins in the VN sensory epithelium of garter<br />

snakes. In addition, by screening a snake VN cDNA library, we<br />

obtained several clones which show high homology to RGS2, RGS3<br />

and RGS4. These RGSs may play a role in desensitization of<br />

chemoattractant-induced signal transduction.<br />

Supported by NIDCD Grant #DC03735


33 Poster [ ] Vomeronasal Organ<br />

CHEMOSIGNAL TRANSDUCTION IN THE VOMERONASAL<br />

ORGAN OF GARTER SNAKES: CHEMOATTRACTANT-<br />

INDUCED MEMBRANE POTENTIAL CHANGES<br />

Liu W. 1, Dalton W. 2, Chen P. 2, Halpern M. 3, Cinelli A. 1 1Anatomy and<br />

Cell Biology, Downstate Medical Center, Brooklyn, NY; 2Biochemistry,<br />

Downstate Medical Center, Brooklyn, NY; 3Anatomy & Cell Biology,<br />

Downstate Medical Center, Brooklyn, NY<br />

We have demonstrated previously that in the snake vomeronasal<br />

(VN) organ chemoattractant (ESS) produce transient calcium cytosolic<br />

accumulation in the dendritic regions of VN neurons via two pathways:<br />

calcium release from IP3-sensitive stores and a plasma membrane<br />

calcium influx. Using voltage-sensitive dyes, here we characterize the<br />

functional role of these events during VN chemosensory transduction.<br />

ESS evokes optical depolarizations which exhibit a time course similar<br />

to the EVG, but shorter than VN calcium transients. Peak<br />

depolarization in the apical VN region are reduced but not suppressed<br />

in the absence of external calcium, indicating that they probably depend<br />

on non-selective cation channels. Depletion of all internal calcium<br />

stores evokes a reduction of these depolarizations, which is not<br />

observed when responses are evoked by potassium depolarization.<br />

Interestingly, the depletion of ryanodine-sensitive calcium stores fails to<br />

evoke this reduction, and instead increases the duration of<br />

depolarizations in the cell body region. This effect is suppressed when<br />

potassium currents are blocked. These results support the notion of a<br />

functional compartmentalization between different calcium stores and<br />

indicates a dual and opposite action of calcium release in snake VN<br />

neurons during chemosensory transduction. Calcium release by IP3sensitive<br />

stores appears to enhance the initial dendritic depolarization,<br />

while calcium release by ryanodine stores seems to activate repolarizing<br />

potassium currents controlling the duration of these responses.<br />

Supported by NIDCD Grant #DC03735<br />

34 Poster [ ] Vomeronasal Organ<br />

VOMERONASAL ORGAN SIGNAL TRANSDUCTION IN THE<br />

CHILEAN LIZARD, LIOLAEMUS BELLII.<br />

Labra A.L. 1, Fadool D.A. 1 1Prog. in Neurosci. & Mol. Biophysics,<br />

Florida State University, Tallahassee, FL<br />

Specimens of Liolaemus bellii (Chilean lizard) were live-captured in<br />

the mountains of central Chile and transported to the U.S.A. to<br />

determine their suitability <strong>for</strong> cellular studies of the vomeronasal organ<br />

(VNO). Rhodamine-conjugated dextran plus 2% Triton X-100 was<br />

introduced into the vomeronasal (VN) orifice to confirm the location of<br />

the VN epithelium and its degree of compartmentalization from the<br />

MOE. After migration of the vital dye <strong>for</strong> two weeks, individual<br />

vomeronasal sensory neurons (VSNs) were distinctly labeled and<br />

showed a bipolar morphology as reported in the VNO of all other<br />

vertebrates. Ten µm cryosections of the VN epithelium were intensely<br />

labeled with antibodies against Giα1-3 and Gβ G-proteins. Extracts<br />

were prepared (1:300 dilution) from three body source secretions (skin,<br />

feces, cloacal) known to contain pheromones mediating chemical<br />

communication in these animals. Three mM agmatine (AGB) was<br />

mixed with one of the body source secretions or control saline and<br />

lizards were stimulated with the pheromone plus AGB mixture to<br />

histologically map and determine secretion rank-order effectiveness.<br />

Using L-cysteine-activated papain <strong>for</strong> VSN isolation, we were able to<br />

preserve both voltage and chemosignal-activated conductances (n=69<br />

total recordings) using recording and pipette solutions optimized <strong>for</strong><br />

turtle preparations. The tuning and percentage of chemosignal-activated<br />

responses (25 of 46 VSNs, 54% response rate), when presented a panel<br />

of five chemical extracts, will be favorable <strong>for</strong> future single cell<br />

electrophysiology studies combined with a distinct cadre of quantifiable<br />

behavioral displays such as the tongue flick, head bob, and scent<br />

marking. Supported by NSF WISE grant.<br />

9<br />

35 Poster [ ] Vomeronasal Organ<br />

PROTEIN INTERACTIONS WITH THE TRPC2 ION CHANNEL<br />

IN THE VOMERONASAL ORGAN (VNO).<br />

Brann J.H. 1, Fadool D.A. 1 1Prog. in Neuroscience & Mol. Biophysics,<br />

Florida State University, Tallahassee, FL<br />

The role of the inositol 1,4,5-trisphosphate (IP )second messenger<br />

3<br />

system in vomeronasal sensory neurons (VSNs) of the vomeronasal<br />

organ (VNO) is unclear. Furthermore, how the functional connections<br />

between the type 3 IP receptor (IP R3) and transient receptor potential<br />

3 3<br />

channel type 2 (TRPC2) may influence the cationic receptor potential is<br />

also unknown. We have previously demonstrated that IP R3 and<br />

3<br />

TRPC2 have an overlapping expression pattern in rodent VNO and<br />

participate in a protein-protein interaction. Here we sought to<br />

demonstrate a functional role <strong>for</strong> the IP R3/TRPC2 protein complex.<br />

3<br />

Pheromone-evoked whole-cell currents were blocked in 4 of 5 VSNs<br />

when a peptide, directed against the interaction domain between<br />

IP R3/TRPC2, was included in the recording pipette held (Vh) at –60<br />

3<br />

mV. Under our recording conditions, pipette dialysis of 240 mM IP3 failed to evoke whole-cell current in 20 of 20 VSNs tested. Typical<br />

antagonists of the IP R (ruthenium red, 2-APB) also failed to block<br />

3<br />

pheromone-evoked currents. SDS-PAGE and Western analysis of male<br />

and female VNO tissue reveals expression of one or more of the long<br />

iso<strong>for</strong>ms of the adaptor protein Homer (1b, 1c, 2a, 2b, and 3; 45 kDa)<br />

as well as neuronal Shc (66, 53, 46 kDa), an adaptor protein known to<br />

communicate G protein-coupled receptor (GPCR) and receptor tyrosine<br />

kinase (RTK) activation. The IP R3/TRPC2 protein complex may be<br />

3<br />

associated within a larger protein scaffold whereby IP and/or adaptor<br />

3<br />

proteins could subserve regulatory functions of the primary<br />

transduction current. Supported by F31 DC06153.<br />

36 Poster [ ] Vomeronasal Organ<br />

SPECIES SPECIFICITY IN RODENT PHEROMONE<br />

RECEPTOR REPERTOIRES<br />

Lane R.P. 1, Young J. 2, Newman T. 2, Trask B.J. 2 1Molecular Biology<br />

and Biochemistry, Wesleyan University, Middletown, CT; 2Division of<br />

Human Biology, Fred Hutchinson Cancer Research Center, Seattle, WA<br />

The mouse V1R putative pheromone receptor gene family consists of<br />

at least 137 intact genes clustered at multiple chromosomal locations in<br />

the genome. Species-specific pheromone receptor repertoires may<br />

partly explain species-specific social behavior. We conducted a<br />

genomic analysis of an orthologous pair of mouse and rat V1R gene<br />

clusters to test <strong>for</strong> species-specificity in rodent pheromone systems.<br />

Mouse and rat have lineage-specific V1R repertoires in each of three<br />

major subfamilies at these loci as a result of post-speciation<br />

duplications, gene loss, and gene conversions. The onset of this<br />

diversification roughly coincides with a wave of Line1 (L1)<br />

retrotranspositions into the two loci. We propose that L1 activity has<br />

facilitated post-speciation V1R duplications and gene conversions. In<br />

addition, we find extensive homology among putative V1R promoter<br />

regions in both species. We propose a regulatory model in which<br />

promoter homogenization could ensure that V1R genes are equally<br />

competitive <strong>for</strong> a limiting transcriptional structure to account <strong>for</strong><br />

mutually exclusive V1R expression in vomeronasal neurons.


37 Poster [ ] Vomeronasal Organ<br />

EXPRESSION PATTERN OF GENES FOR NOTCH<br />

SIGNALING PATHWAY IN MOUSE VOMERONASAL ORGAN<br />

DURING ONTOGENY AND REGENERATION AFTER<br />

REMOVAL OF ACCESSORY OLFACTORY BULB.<br />

Wakabayashi Y. 1, Ichikawa M. 2 1Research Fellow of the Japan Society<br />

<strong>for</strong> the Promotion of Science, Tokyo, Japan; 2Dep Basic Tech and<br />

Facilities, Tokyo Metropol Inst Neurosci, Tokyo, Japan<br />

Vomeronasal receptor neurons (VRNs) proliferate and differentiate<br />

continuously throughout life. Proliferation of VRNs mainly occurs in<br />

the marginal region of the sensory epithelium of adult vomeronasal<br />

organ (VNO). The Notch signaling pathway is involved in cell fate<br />

decisions and differentiation during developmental CNS. We have<br />

studied whether Notch signaling pathway involves in differentiation<br />

and proliferation of VRNs during ontogeny and regeneration.<br />

In this study, we examined the expression patterns of Notch and<br />

their ligands, Delta and Jagged, using in situ hybridization during<br />

ontogeny and regeneration after removal of accessory olfactory bulb<br />

(AOBX) in mice. In adult VNO, a few Notch1(+), Delta1(+) and<br />

BrdU(+) cells appeared only in the marginal region, whereas Jagged2<br />

was expressed in all VRNs. Notch1(+), Delta1(+) and BrdU(+) cells<br />

located close to the basal lamia at embryonic days 14.5 and 16.5,<br />

whereas expression level of Jagged2 was very low in comparison with<br />

adult. At AOBX day2, number and location of Notch1(+), Delta1(+)<br />

and BrdU(+) cells did not change. Expression pattern of Jagged2 did<br />

not change. At AOBX day7, large amount of Notch1(+), Delta1(+) and<br />

BrdU(+) cells appeared in the marginal region, whereas expression<br />

pattern of Jagged2 did not change. These results suggested that the<br />

interaction of Notch1(+) and Delta1(+) cells play important roles in<br />

vomeronasal neurogenesis of VNO during ontogeny as well as<br />

regeneration after AOBX mice.<br />

38 Poster [ ] Vomeronasal Organ<br />

THE ROLE OF THE VOMERONASAL SYSTEM IN FOOD<br />

PREFERENCES OF THE GRAY SHORT-TAILED OPOSSUM,<br />

MONODELPHIS DOMESTICA<br />

Daniels Y. 1, Halpern M. 2, Zuri I. 1 1Anatomy and Cell Biology,<br />

Downstate Medical Center, Brooklyn, NY; 2Anatomy & Cell Biology,<br />

Downstate Medical Center, Brooklyn, NY<br />

The vomeronasal system (VNS) is usually considered primarily a<br />

pheromone-detecting system. In snakes and lizards, this system is also<br />

important <strong>for</strong> feeding behavior. To date, no studies have reported<br />

feeding deficits in mammals deprived of a functional VNS. M.<br />

domestica is considered a primitive mammalian species that was<br />

recently introduced into laboratories. Since these opossums respond to a<br />

variety of foods, they are a good model to investigate the role of the<br />

VNS in food preferences in mammals.<br />

The six male and seven female gray short-tailed opossums used in<br />

this study were simultaneously presented with four foods, one from<br />

each of the following food groups: fruits (apples, oranges, peaches,<br />

cantaloupes), meats (mealworms, chicken, pork, crickets), processed<br />

vegetables (raisin bran, cheerios, whole wheat bread, bagel) and<br />

unprocessed vegetables (corn, peppers, carrots, broccoli). Be<strong>for</strong>e<br />

blocking access to the vomeronasal organ (VNO) with gel foam and<br />

Crazy Glue, the opossums selected meats most frequently and fruits<br />

more frequently than processed and unprocessed vegetables. Following<br />

VNO blockage, the opossums demonstrated no preference between the<br />

different food groups. This study suggests that without a functional<br />

vomeronasal organ, the food preferences of gray short-tailed opossums<br />

are significantly impaired.<br />

Supported by NIDCD Grant # DC02745.<br />

10<br />

39 Slide [ ] Vomeronasal Organ<br />

CHEMO-INVESTIGATORY BEHAVIOUR OF MALE MICE IN<br />

DETECTING ESTRUS: ROLE OF OLFACTORY-<br />

VOMERONASAL SYSTEM<br />

Raman S.A. 1 1Animal Scinece, Bharathidasan University,<br />

Trichirappalli, India<br />

S.ACHIRAMAN AND G.ARCHUNAN<br />

Department of Animal Science, Bharathidasan University<br />

Tiruchirappalli-620 024, Tamilnadu.<br />

INDIA<br />

achiraman_s@yahoo.co.in<br />

The aim of the present study is to evaluate whether the male<br />

mouse is capable of discriminating the female urinary odour of different<br />

reproductive phases (with a view to detect estrus using Y-maze<br />

apparatus and to establish the relationship of olfactory - vomeronasal<br />

system and chemo-investigatory behaviour in estrus detection. Hence,<br />

normal, vomeronasal organ (VNO)-ablated and Zinc Sulphate-irrigated<br />

mouse were used as test animals. Various behaviours such as frequency<br />

of visit, duration of visit, sniffing, licking, body rubbing and grooming<br />

were recorded. The normal mice frequently visited and devoted more<br />

time in delivering various behaviours towards the estrus urine sample in<br />

comparison to that of non-estrus urine. The VNO-ablated mice showed<br />

significant reduction in response to estrus urine (duration of visit and<br />

self-grooming) than that of ZnSO4-irrigated mice. However, the<br />

ZnSO4-irrigated mice showed significant reduction in frequency of<br />

visits to the urine samples. These results clearly reveal that the VNO<br />

play a significant role in the detection of estrus in mice. The present<br />

results also suggest that male mice preferentially communicate sexual<br />

interest via self-grooming towards the opposite sex. By self-grooming<br />

at higher rates, male mice may be broadcasting scents to attract<br />

potential mates or to in<strong>for</strong>m their willingness to mate.<br />

40 Poster [ ] Vomeronasal Organ<br />

NEUROGENESIS, MIGRATION AND APOPTOSIS IN THE<br />

VOMERONASAL EPITHELIUM OF ADULT MICE<br />

Martinez-Marcos A. 1, Quan W. 2, Jia C. 2, Halpern M. 3 1Departamento<br />

de Ciencias Medicas, Universidad de Castilla-La Mancha, Albacete,<br />

Albacete, Spain; 2Anatomy and Cell Biology, Downstate Medical<br />

Center, Brooklyn, NY; 3Anatomy & Cell Biology, Downstate Medical<br />

Center, Brooklyn, NY<br />

Neurogenesis in the adult mouse vomeronasal organ appears to occur<br />

in the central regions, but is more prevalent at the edges of sensory<br />

epithelium. Basal cells at the center of the epithelium participate in cell<br />

replacement. It is unknown whether dividing cells at the edges<br />

constitute a reservoir <strong>for</strong> growth, become apoptotic or participate in<br />

neural turnover. This latter possibility implies a process of horizontal<br />

migration. The present work addresses this controversy by injecting<br />

bromodeoxyuridine in adult mice and allowing them to survive <strong>for</strong><br />

various intervals. The vertical and horizontal position of labeled cells<br />

was analyzed as a function of time. Both, vertical and horizontal<br />

migration of labeled cells were detected. Cells at the center of the<br />

epithelium migrate vertically to become neurons as demonstrated by coexpression<br />

of olfactory marker protein. Cells at the edges migrate<br />

horizontally toward the center. After 42 days, however, they have<br />

migrated less than 10% of the distance from the edge (0%) to the center<br />

of the epithelium (100%), thus making it likely that if these cells<br />

participate in neural turnover it is only in marginal regions. The pattern<br />

of distribution of apoptotic cells has been studied and, interestingly, it is<br />

similar to that of dividing cells. These results support the idea that<br />

sensory cell renewal in the mouse vomeronasal organ occurs through a<br />

process of vertical migration. Supported by grant DC02745.


41 Poster [ ] Vomeronasal Organ<br />

NEW METHOD OF VOMERONASAL NERVE TRANSECTION<br />

LEAVES THE OLFACTORY SYSTEM INTACT.<br />

Matsuoka M. 1, Norita M. 2, Costanzo R.M. 3 1Niigata University<br />

Graduate School of Medical and Dental <strong>Sciences</strong>, Niigata, Niigata,<br />

Japan; 2Division of Neurobiol. & Anat., Niigata University Graduate<br />

School of Medical and Dental <strong>Sciences</strong>, Niigata, Niigata, Japan;<br />

3Physiology, Virginia Commonwealth University, Richmond, VA<br />

Methods used to study degeneration and regeneration in the<br />

vomeronasal system typically result in collateral damage to olfactory<br />

system pathways. We set out to develop a new approach to selectively<br />

lesion the vomeronasal nerves and to leave the olfactory system and its<br />

nerve fibers intact. For this study we used OMP-tauLacZ mice and Xgal<br />

staining, which allowed us to label both vomeronasal and olfactory<br />

pathways containing the Olfactory Marker Protein (OMP). Vannis<br />

micro-dissecting scissors were used to cut the vomeronasal nerves at a<br />

point just anterior to the Accessory Olfactory Bulb (AOB). We then<br />

observed the vomeronasal epithelium, the olfactory epithelium, the<br />

main olfactory bulb (MOB) and the AOB at 1, 6, 20, 60 and 120 days<br />

after nerve lesion. At days 20 and 60 we found that there were no OMP<br />

positive cells in the vomeronasal epithelium and no OMP positive nerve<br />

terminals in the AOB. In contrast, OMP positive cells in the olfactory<br />

epithelium and their OMP positive nerve fiber projections to the MOB<br />

remained intact. In a few animals examined after 120 days of recovery<br />

we noticed OMP staining in both the vomeronasal epithelium and<br />

nerve projections within the AOB. These findings suggest that recovery<br />

in the vomeronasal epithelium may be associated with the retargeting of<br />

nerve fibers in the AOB. This new approach to vomeronasal nerve<br />

transaction may prove important <strong>for</strong> behavioral studies of vomeronasal<br />

function, since it has the distinct advantage of leaving the olfactory<br />

system intact. Supported by NIH grant DC00165.<br />

42 Slide [ ] Vomeronasal Organ<br />

PROTEOMICS ANALYSIS OF OLFACTORY SYSTEMS: A<br />

GENERIC APPROACH USING DIFFERENTIAL DYE<br />

ANALYSIS AND DE NOVO PEPTIDE SEQUENCING FOR<br />

IDENTIFYING PROTEINS POTENTIALLY INVOLVED IN<br />

PHEROMONE TRANSPORT AND RECEPTION.<br />

Greenwood D. 1, Jordan M. 1, Cooney J. 2, Jensen D. 2, Plummer K. 3,<br />

Newcomb R. 1, Rasmussen L. 4 1Gene Technologies, HortResearch,<br />

Auckland, New Zealand; 2HortResearch, Hamilton, New Zealand;<br />

3School of Biological <strong>Sciences</strong>, University of Auckland, Auckland, New<br />

Zealand; 4Oregon Graduate Institute of Science & Technology,<br />

Beaverton, OR<br />

Mucosal proteins from the trunk and VNO duct of the Asian elephant<br />

and proteins, both soluble and membrane-bound, extracted from the<br />

antennae of tortricid moths have provided us with diverse sources of<br />

olfactory proteins. We have developed a proteomics plat<strong>for</strong>m which<br />

concentrates on targeting proteins that are differentially expressed, such<br />

as exhibiting sex-specific expression. This approach enables us to<br />

significantly reduce the total number of proteins that feed into our<br />

proteomics pipeline. Proteins conjugated from reaction with<br />

multiplexed fluorescent dyes are separated on 1-D and 2-D<br />

polyacrylamide gels and scanned at segregating wavelengths. The<br />

resultant images are processed using gel analysis software. Those<br />

proteins displaying significantly altered levels or in a new position are<br />

excised and subjected to trypsin hydrolysis and released peptides are<br />

analyzed by nanoelectrospray ion-trap mass spectrometry. Primary data<br />

crunching is per<strong>for</strong>med using peptide search algorithms interrogating<br />

both public domain and proprietary databases. Emphasis is placed on<br />

data generated from de novo sequencing of peptides using the<br />

DeNovoX® package followed by BLAST analysis in an attempt to<br />

overcome the shortfall of having incompletely sequenced genomes.<br />

11<br />

43 Poster [ ] Olfactory Transduction<br />

G-PROTEINS IN ANOPHELES GAMBIAE OLFACTION<br />

Rützler M. 1, Zwiebel L. 1 1Biological <strong>Sciences</strong>, Vanderbilt University,<br />

Nashville, TN<br />

<strong>Chemoreception</strong> in general and olfaction in particular represent<br />

critical sensory inputs into many behaviors of hematophagous (bloodfeeding)<br />

insects. Recently the first odorant receptors (ORs) of the<br />

malaria vector mosquito Anopheles gambiae (AgORs) have been<br />

identified and functionally characterized. In addition to these studies, it<br />

will be important to develop methodology to facilitate high throughput<br />

screening of substances, which could potentially lead to the<br />

development of a new generation of efficient insect repellents. To build<br />

an experimental setup that could allow such an approach, expansion of<br />

our current knowledge about signal transduction in insect olfaction is<br />

crucial. The current study identifies 6 genes in the Anopheles gambiae<br />

genome that encode G-protein (α-subunit) homologues. We have<br />

investigated tissue-specific expression of these Gα-genes and describe<br />

the expression pattern of a total of 11 transcripts in the adult mosquito.<br />

Furthermore we localized Gα-proteins within the female mosquitoantenna<br />

and identify candidates <strong>for</strong> Gα-proteins involved in olfactory<br />

signal transduction. Supported by grants from NIAID and NIDCD.<br />

44 Poster [ ] Olfactory Transduction<br />

THE NOVEL GUANYLYL CYCLASE MSGC-I MAY MEDIATE<br />

PHEROMONE-INDUCED CGMP IN THE ANTENNAE OF<br />

MANDUCA SEXTA<br />

Fernando S. 1, Collmann C. 2, Nighorn A. 1 1ARLDN, University of<br />

Arizona, Tucson, AZ; 2Neurobiology, University of Arizona, Tucson, AZ<br />

The signal transduction cascade induced by pheromone detection<br />

results in a transient increase in calcium followed by a slow increase in<br />

cGMP, which is thought to play a role in adaptation in the olfactory<br />

receptor neurons (ORNs) of Manduca sexta. However, the specific<br />

mechanism mediating this increase in cGMP is unknown. We wanted to<br />

determine if the novel guanylyl cyclase, MsGC-I, mediates this increase<br />

in cGMP. Using immunocytochemistry and in situ hybridization, we<br />

found that MsGC-I is likely to be expressed in at least a subset of<br />

pheromone-sensitive ORNs. Some neuronal calcium sensor (NCS)<br />

proteins are known to regulate calcium-dependent guanylyl cyclase<br />

activity in the olfactory system. To learn how the activity of MsGC-I<br />

might be regulated we are testing if MsFrequenin and/or<br />

MsNeurocalcin, two NCSs cloned previously and known to be<br />

expressed in ORNs, may mediate calcium-dependent regulation of<br />

MsGC-I activity. Supported by NIH-NIDCD DC04292


45 Slide [ ] Olfactory Transduction<br />

EVIDENCE FOR ALTERNATE TRANSDUCTION PATHWAYS<br />

IN THE MOUSE: OLFACTION IN THE CNGA2 KNOCKOUT<br />

Restrepo D. 1, Lin W. 1, Arellano J. 1, Slotnick B. 2 1Cell and<br />

Developmental Biology, Neuroscience Program and Rocky Mountain<br />

Taste and Smell Center, University of Colorado Health <strong>Sciences</strong><br />

Center, Denver, CO; 2Psychology, University of South Florida, Tampa,<br />

FL<br />

In mammals the adenosine 3´,5´-monophosphate (cAMP) pathway is<br />

widely believed to be the only transduction mechanism in the main<br />

olfactory epithelium largely due to the lack of odor responsiveness of<br />

mice defective <strong>for</strong> cAMP signaling. Here we report on odor<br />

responsiveness in mice with a disrupted cyclic nucleotide gated (CNG)<br />

channel subunit A2. Several odorants, including putative pheromones,<br />

can be detected and discriminated by these mice behaviorally. These<br />

odors elicit electro-olfactogram (EOG) responses in the olfactory<br />

epithelium and activate a subset of glomeruli in the main olfactory bulb<br />

and stimulate neurons in of CNGA2 knockout mice. In addition, EOG<br />

responses to odors detected by CNGA2 knockout mice are relatively<br />

insensitive to inhibitors of the cAMP pathway. These results provide<br />

strong evidence that cAMP-independent pathways in the main olfactory<br />

system of mammals participate in detecting a subset of odors.<br />

This work was supported by NIH grants DC00566, DC04657,<br />

DC006070 (DR) DC0043 (WL) and MH6118 (BS).<br />

46 Slide [ ] Olfactory Transduction<br />

IDENTIFICATION OF THE SECOND MESSENGER THAT<br />

MEDIATES SIGNAL TRANSDUCTION IN THE NEWT<br />

OLFACTORY RECEPTOR CELL<br />

Takeuchi H. 1, Kurahashi T. 1 1Frontier Biosciences, Osaka University,<br />

Osaka, Japan<br />

It has long been believed that vertebrate olfactory signal transduction<br />

is mediated by independent multiple pathways (utilizing cAMP and IP3<br />

as second messengers). However, the dual presence of parallel<br />

pathways in the olfactory cilia is still controversial, mainly because of<br />

the lack of in<strong>for</strong>mation regarding the single cell response induced by<br />

odorants that have been shown to produce IP3 exclusively.<br />

In the present study, we examined two series of experiments. First,<br />

we compared responses induced by both cAMP- and IP3-odorants.<br />

Fundamental properties of responses were surprisingly homologous in<br />

spatial distribution of the sensitivity, wave<strong>for</strong>ms, I-V relation and<br />

reversal potential, dose-dependence, time integration of stimulus,<br />

adaptation and recovery. By applying both types of odorants<br />

alternatively to the same cell, we observed cells to exhibit perfect<br />

symmetrical cross-adaptation. Second series of experiments was that<br />

the cytoplasmic cNMP concentration was manipulated through the<br />

photolysis of caged compounds to examine their real-time interactions<br />

with IP3-odorant-induced events. The Properties of responses induced<br />

by both IP3-odorants and cytoplasmic cNMP resembled each other in<br />

their unique characteristics. They showed symmetrical adaptation that is<br />

dependent on the Ca2+-infux. Furthermore, both responses were<br />

additive in a manner as predicted quantitatively by the theory that signal<br />

transduction is mediated by the increase in cytoplasmic cAMP. The<br />

data will provide evidence showing that olfactory responses are<br />

generated by a uni<strong>for</strong>m mechanism <strong>for</strong> a wide variety of odorants.<br />

12<br />

47 Poster [ ] Olfactory Transduction<br />

PHARMACOLOGICAL PROPERTIES OF A POSSIBLE TRP-<br />

RELATED ION CHANNEL IN LOBSTER OLFACTORY<br />

RECEPTOR NEURONS<br />

Bobkov Y.V. 1, Ache B.W. 1 1The Whitney Laboratory, Center <strong>for</strong> Smell<br />

and Taste, McKnight Brain Institute, University of Florida, Gainesville,<br />

FL<br />

Lobster olfactory receptor neurons express a sodium-gated nonselective<br />

cation channel (Zhainazarov et al., J. Neurophysiol.79:1349,<br />

1998) that is a potential member of the growing family of trp channels.<br />

Here, we extend our pharmacological characterization of the channel by<br />

recording the effect of potential agonists and antagonists on the channel<br />

in cell-free patches. In addition to Na + , the channel is activated by Ca2+ ,<br />

PIP2, PIP3, and maitotoxin, and antagonized by H + , calmodulin<br />

inhibitors, and the trp channel blockers 2APB, SKF96365, ruthenium<br />

red, Al3+ , Gd3+ , and La3+ . Interestingly, decreasing extracellular pH<br />

from 8.0 to 7.5, a range in which pH is not known to block other ion<br />

channels, effectively blocked the lobster channel and could serve as a<br />

selective probe <strong>for</strong> the channel. We are in the process of using this<br />

enhanced pharmacological profile to implicate the channel as a target of<br />

odor-activated phosphoinositide signaling in the cells. Supported by the<br />

NIDCD (DC01655).<br />

48 Poster [ ] Olfactory Transduction<br />

EXPRESSION OF TRP CHANNELS AND ODORANT<br />

RECEPTOR SIGNALING IN ODORA CELLS<br />

Liu G. 1, Talamo B.R. 1 1Neuroscience, Tufts University, Boston, MA<br />

To examine signaling pathways directly activated by odorant<br />

receptor, we utilize the olfactory epithelial cell line, Odora. Odora cells<br />

functionally express transfected odorant receptor U131 without<br />

requiring concurrent transfection of additional signaling proteins<br />

(Murrell and Hunter, 1999). Previously we showed that Odora cells<br />

express adenylyl cyclase III, phospholipase C iso<strong>for</strong>ms, and G proteins<br />

found in the olfactory epithelium (OE). Cai signals in Odora cells are<br />

not activated by the canonical cAMP cascade; transfected U131<br />

receptors in Odora cells activate PLC, IP3 and IP3 receptor to elevate<br />

Cai without the release of intracellular Ca2+. Biotinylation studies<br />

reveal that there is cell surface expression of IP3 receptors I-III in<br />

Odora cells, implicating one or more IP3 receptors as a possible route<br />

<strong>for</strong> Ca2+ entry mediated by odorant. In this study, we explore the<br />

expression of transient receptor potential (TRP) channels to assess the<br />

possibility that they might participate in odor-activated Ca2+ entry<br />

through the plasma membrane. RT-PCR analysis of all 7 mammalian<br />

TRPC channels reveals transcripts <strong>for</strong> TRPC1 and TRPC5 in both<br />

Odora cells and nasal epithelium. Western blotting further documents<br />

the expression of protein <strong>for</strong> TRPC5, but not <strong>for</strong> TRPC1. Current<br />

studies using antibodies to TRPC5 and TRPC1 examine the localization<br />

of TRPCs in the rat olfactory system. These observations continue the<br />

characterization and evaluation of a unique pathway activated by<br />

odorant receptor in Odora cells that may reveal an important role <strong>for</strong><br />

alternative signaling through the odorant receptor in the olfactory<br />

system in vivo. Supported in part by DC 05229-01A1.


49 Poster [ ] Olfactory Transduction<br />

ODOR-INDUCED CALCIUM RESPONSES FROM SQUID<br />

OLFACTORY RECEPTOR NEURONS<br />

Sitthichai A.A. 1, Lucero M.T. 1 1Physiology, University of Utah, Salt<br />

Lake City, UT<br />

ORNs from the squid Lolliguncula brevis fall into 5 morphological<br />

subtypes. Previously we characterized odor responses in 2<br />

morphological subtypes from isolated squid olfactory receptor neurons<br />

(ORNs) 1 . Because the isolation process destroys the other 3 subtypes,<br />

we have developed a slice preparation to make physiological recordings<br />

from a relatively intact olfactory epithelium. Thus far,<br />

electrophysiological recordings from slices have been challenging<br />

because of the thick mucous that covers the olfactory organ. Initial<br />

Ca2+ imaging studies using the fluorescent Ca2+ indicator dye, Fluo-4,<br />

and laser scanning confocal microscopy were unsuccessful because the<br />

tissue underlying the olfactory epithelium spontaneously contracts after<br />

dissection from the animal. We tested numerous receptor antagonists,<br />

ion channel inhibitors and conotoxins that were unsuccessful in<br />

blocking spontaneous contractions. Both isotonic MgCl and 62.5<br />

2<br />

µg/ml ketamine blocked the tissue contractions but also inhibited odorinduced<br />

Ca2+ responses. Fortunately, we found that 5 µM nicotine<br />

completely blocks spontaneous contractions without compromising<br />

odor responses. We are currently using the squid olfactory organ slice<br />

preparation to test the odor specificity and transduction pathway of the<br />

5 morphological subtypes of ORNs. Application of this methodology<br />

will allow us to correlate morphology, odor specificity and signal<br />

transduction in real time across a relatively intact olfactory epithelium.<br />

Funded by NINDS NS07938 (MTL) and NIDCD DC006793 (AAS).<br />

1. M. T. Lucero, F. T. Horrigan, and W. F. Gilly. J.Exp.Biol. 162:231-<br />

249, 1992.<br />

50 Slide [ ] Olfactory Transduction<br />

NOVEL ACTION OF CALMODULIN ON NATIVE RAT<br />

OLFACTORY CNG CHANNELS<br />

Bradley J. 1, Bönigk W. 2, Gensch T. 2, Yau K. 1, Kaupp B. 2, Frings S. 3<br />

1Department of Neuroscience/HHMI, Johns Hopkins University,<br />

Baltimore, MD; 2IBI-1, Forschungszentrum, Jülich, Germany;<br />

3Molecular Physiology, University of Heidelberg, Heidelberg, Germany<br />

Adaptation to stimuli is an important process in sensory neurons. In<br />

vertebrate olfactory receptor neurons, rapid negative-feedback<br />

inhibition of Ca2+-permeable, cAMP-gated (CNG) transduction<br />

channels underlies adaptation. Previous studies with CNGA2, a CNG<br />

channel subunit that <strong>for</strong>ms a homomeric channel when expressed<br />

heterologously, have implicated binding of Ca2+/calmodulin<br />

(Ca2+/CaM) to CNGA2 in adaptation. Native rat olfactory CNG<br />

channels, however, are a heteromeric complex of three homologous<br />

subunits, CNGA2, CNGA4 and CNGB1b. We now report that the<br />

CaM-binding site on CNGA2 does not mediate at all the modulation by<br />

Ca2+-CaM of native channels. We find that, with native channels in<br />

resting neurons, Ca2+-free calmodulin (apocalmodulin) is largely preassociated.<br />

Accordingly, apocalmodulin-binding sites of the IQ-type are<br />

together necessary and sufficient <strong>for</strong> Ca2+/CaM feedback inhibition of<br />

heteromeric channels. Thus, calmodulin is permanently associated with<br />

native channels in ORNs, poised at the site of Ca2+ influx as a Ca2+<br />

sensor to rapidly effect inhibition. Apart from mechanistically<br />

explaining the fast kinetics of adaptation to odorants, our findings<br />

caution against the practice of extrapolating findings from<br />

heterologously expressed homomeric channels to native heteromeric<br />

channels.<br />

13<br />

51 Slide [ ] Olfactory Transduction<br />

AMPLIFICATION BY A SINGLE GPCR MOLECULE IN THE<br />

OLFACTORY RECEPTOR NEURON<br />

Bhandawat V. 1, Reisert J. 2, Yau K. 3 1Neuroscience, Johns Hopkins<br />

University, Baltimore, MD; 2Howard Hughes Medical Institute,<br />

Baltimore, MD; 3Neuroscience (HHMI), Johns Hopkins University,<br />

Baltimore, MD<br />

Odorants bind to specific G-protein coupled receptors (GPCRs) on<br />

olfactory neurons, which, via a G-protein cascade, convert chemical<br />

signals into an electrical response. We have measured membrane<br />

current from single frog (Rana pipiens) olfactory neurons and analyzed<br />

the odorant-receptor interaction. We find that an odorant molecule<br />

rapidly unbinds from the receptor, often be<strong>for</strong>e the activation of a Gprotein.<br />

This leads to low amplification between receptor activation and<br />

G-protein activation, which is confirmed by the small size of the unitary<br />

event as derived by quantal analysis. Although the individual events are<br />

very small, there is a non-linear summation of these events that<br />

amplifies the response greatly. Funded by Howard Hughes Medical<br />

Institute.<br />

52 Slide [ ] Olfaction: Animal Behavior<br />

A COMPARISON OF SENSORY HAIR DISTRIBUTION ON<br />

THE CHELAE AND OLFACTORY ORGANS OF CRAYFISH<br />

(ORCONECTES RUSTICUS)<br />

Bergman D.A. 1, Belanger R.M. 1, Moore P.A. 1 1Biological <strong>Sciences</strong>,<br />

Bowling Green State University, Bowling Green, OH<br />

Sensory hairs are found in great abundance on the appendages of<br />

crayfish and have been shown to be important <strong>for</strong> detecting both<br />

mechano- and chemosensory stimuli. Detailed analysis of the spatial<br />

distribution of these sensory hairs that is correlated with sex or<br />

reproductive <strong>for</strong>ms allows us to determine if a particular sensory hair<br />

type or abundance could be important <strong>for</strong> mating. With this in mind, we<br />

examined the sensory hair distribution of <strong>for</strong>m I (reproductive) and<br />

<strong>for</strong>m II (non-reproductive) males, as well as females using scanning<br />

electron microscopy. We quantified and qualified all sensory hair types.<br />

Our research indicates that reproductive male crayfish have a greater<br />

number of feathered hairs on their chelae and antennules when<br />

compared to non-reproductive males and females. This increase in<br />

sensory hair distribution suggests a link between sensory hairs and<br />

mating chemical cues in the crayfish. Male and female crayfish have<br />

different distributions and abundance of sensory hairs on the sensory<br />

appendages. These differences are likely a function of the different<br />

reproductive life histories where males actively pursue and locate<br />

females. In all, the sensory hairs on all of the appendages appear to be<br />

important in mate recognition.


53 Poster [ ] Olfaction: Animal Behavior<br />

AESTHETASCS, THE OLFACTORY SENSILLA, ARE<br />

MEDIATORS OF CHEMOSENSORY ACTIVATION OF<br />

ANTENNULAR FLICKING IN THE SPINY LOBSTER,<br />

PANULIRUS ARGUS<br />

Daniel P.C. 1 1Biology, Hofstra University, Hempstead, NY<br />

Lobster antennules bear a number of different sensilla sensitive to<br />

odorants. We have found that two antennular behaviors, grooming<br />

(Wroblewska et al., Chem. Senses 27:769-778, 2002) and flicking (Fox<br />

et al. Chem. Senses 28:555, 2003), are mediated by aesthetascs and/or<br />

asymmetric setae located in the “tuft” region of the lateral flagellum.<br />

Flicking also requires “non-tuft” setae scattered across the lateral and<br />

medial flagella. Schmidt et al. (this meeting) have found through<br />

ablation experiments that grooming is solely activated by asymmetric<br />

setae. The present study used similar techniques to determine the<br />

relative importance of aesthetascs and asymmetric setae to activation of<br />

flicking behavior. Two groups of lobsters were sham-ablated by<br />

removing a row of guard setae and then tested <strong>for</strong> antennular responses<br />

to L-glutamate (the major chemical stimulus eliciting grooming), and<br />

squid extract (concentration). This was followed by either ablation of<br />

asymmetric setae (N=8) or ablation of aesthetascs (N=6) after which the<br />

behavioral assay was repeated. Lobsters with asymmetric setae<br />

removed no longer groomed in response to glutamate but showed no<br />

reduction in flick rate to squid extract compared to their responses to<br />

the same odorants following sham ablation. In contrast, lobsters with<br />

aesthetascs removed showed no reduction in grooming to glutamate but<br />

no longer increased flick rates to squid extract. These results, along<br />

with those from the earlier study, suggest that olfactory and<br />

nonolfactory pathways are utilized in eliciting flicking behavior in<br />

which processing occurs via the olfactory lobes and the lateral<br />

antennular neuropils.<br />

54 Poster [ ] Olfaction: Animal Behavior<br />

BEHAVIORAL DISCRIMINATION OF AMINO ACIDS IN<br />

ZEBRAFISH (DANIO RERIO)<br />

Valentincic T. 1, Miklavc P. 1 1Department of Biology, University of<br />

Ljubljana, Ljubljana, Slovenia<br />

Based on the dissimilar patterns of glomerular activation during<br />

amino acid stimulation in recent calcium-imaging studies in zebrafish<br />

(Friedrich and Korsching, 1997), it is possible, but previously untested,<br />

that this species is capable of discriminating behaviorally different<br />

amino acids. As per<strong>for</strong>med in catfish, olfactory discrimination was<br />

studied by conditioning zebrafish with a food reward that was presented<br />

90 seconds after the conditioning amino acid solution was injected into<br />

the aquarium. Tests <strong>for</strong> discrimination began after ~50 conditioning<br />

sessions. Conditioned zebrafish, which associated a specific amino acid<br />

odor with a food reward, searched <strong>for</strong> food longer and more intensely<br />

(measurements of the swimming path by video-tracking or counting the<br />

turns >90 degrees during 90 seconds) than after stimulation with a nonconditioned<br />

amino acid. We used 3x10-5M L-Ala, L-Val and L-Arg,<br />

respectively, as conditioning stimuli and the following as test stimuli:<br />

Gly, L-Ala, L-Ser, L-Phe, L-Tyr, L-Trp, L-His, L-Asn, L-Val, L-Ile, L-<br />

Leu, L-Met, L-Arg, L-Lys, L-Glu, L-Asp, L-Pro and D-Ala . With the<br />

exception of the L-Val conditioning stimulus and the L-Ile test<br />

stimulus, zebrafish discriminated all the conditioning stimuli from the<br />

test stimuli. Our results clearly indicated that the amino acids that<br />

previously showed similar glomerular activity patterns (e.g. L-Val and<br />

L-Ile), were not discriminated behaviorally by zebrafish, but those in<br />

which the glomerular activation patterns were distinct were readily<br />

discriminated.<br />

Friedrich, R.W. and Korsching, S., 1997. Combinatorial and<br />

chemotopic odorant coding in the zebrafish olfactory bulb visualised by<br />

optical imaging. Neuron 18:737-752.<br />

14<br />

55 Poster [ ] Olfaction: Animal Behavior<br />

OLFACTORY COMMUNICATION: EVOLVING NEW BLENDS<br />

AND NOVEL PREFERENCES<br />

Vickers N.J. 1, Hillier K. 1, Groot A. 2, Gould F.L. 2 1Biology, University<br />

of Utah, Salt Lake City, UT; 2Entomology, North Carolina State<br />

University, Raleigh, NC<br />

Male moths are often extremely sensitive and narrowly tuned to the<br />

blend of components emitted by conspecific females. The sexual<br />

communication channel is there<strong>for</strong>e under strong stabilizing selective<br />

pressure. How then do new female blends and male preferences evolve?<br />

Two closely related moth species, Heliothis virescens and Heliothis<br />

subflexa, can be hybridized under laboratory conditions in order to<br />

study the genetic basis of behavior and olfactory characteristics that<br />

accompany species divergence. Males of these two species prefer<br />

qualitatively distinct blends that include either Z9-14:Ald (H. virescens)<br />

or Z9-16:Ald (H. subflexa). In addition, H. subflexa males require Z11-<br />

16:OH. These behavioral preferences are correlated with the specificity<br />

of olfactory receptor neurons and central interneurons arborizing within<br />

the glomeruli of the male-specific macroglomerular complex. Wind<br />

tunnel studies have shown that Z9-14:Ald/Z9-16:Ald preference<br />

segregates in a 1:1 ratio in backcross males. Preliminary genetic (QTL)<br />

analyses have revealed that this trait is associated with a single<br />

autosomal chromosome suggesting that a major gene may control this<br />

phenotype, perhaps coupled to the expression of odorant receptors.<br />

Other divergent characters in this system, such as Z11-16:OH-agonism<br />

and Z11-16:OAc-antagonism, may involve odorant receptors but also<br />

likely involve shifts in the glomerular targets of receptor axons and<br />

interpretation of olfactory in<strong>for</strong>mation by higher brain centers.<br />

Supported by NSF IBN-9905683 to NJV.<br />

56 Poster [ ] Olfaction: Animal Behavior<br />

OLFACTORY RECOGNITION IN CANINES<br />

Puchalski D. 1, Leitch A. 1, Hornung D. 1 1Biology Department, St.<br />

Lawrence University, Canton, NY<br />

The overall objective of this work was to assess the specificity of the<br />

olfactory cues that canines use when identifying humans. That is, once<br />

a dog is trained to identify the smell of a human (the target), will the<br />

dog be confused by the smell of the target´s siblings or by the smell of<br />

other people who use the target´s soap or deodorant? A 1-year-old<br />

golden retriever was trained to pick her owner´s scent out of three<br />

possible choices. After a trial had been initiated by “go” the dog would<br />

smell each of three boxes containing T-shirts impregnated with human<br />

scent. The target was randomly assigned to one of the boxes. The boxes<br />

were constructed such that the dog could not use visual cues. The dog<br />

signaled recognition of the target´s smell by a sit/stay response. The<br />

dog was trained to correctly identify the target over 90% of the time.<br />

When there was no target present the dog would repeatedly sample the<br />

three boxes and 80% of the time give no recognition signal. The dog<br />

was rewarded only <strong>for</strong> correct target responses. Probe trials consisting<br />

of the scent of the target´s siblings or of the scent of other humans who<br />

had used the soap and/or deodorant of the target were inserted into the<br />

testing sessions. There was no reward given during a probe trial. The<br />

dog gave the recognition response more often to the smell of the<br />

target´s siblings as compared to the smell of non-related humans.<br />

Likewise the dog gave the recognition response more often when nonrelated<br />

humans used either the soap or deodorant of the target. These<br />

results suggest that both genetic factors (such as HLA typing) and<br />

added smells (such as soap) may have some bearing on canine olfactory<br />

recognition of humans.


57 Poster [ ] Olfaction: Animal Behavior<br />

MLPEST AS A MEASURE OF OLFACTORY SENSITIVITY<br />

THRESHOLD IN MICE<br />

Clevenger A.C. 1, Restrepo D. 1 1Cell and Developmental Biology,<br />

University of Colorado Health <strong>Sciences</strong> Center, Denver, CO<br />

Threshold is defined as the stimulus intensity necessary <strong>for</strong> a subject<br />

to reach a specified percent correct on a detection test. Un<strong>for</strong>tunately,<br />

many measures of threshold in sensory systems are heavily influenced<br />

by decision making processes and are subject to extinction. MLPEST<br />

(Maximum-Likelihood Parameter Estimation by Sequential Testing) is<br />

a method that minimizes both decision-based bias and extinction.<br />

Originally developed <strong>for</strong> human auditory and visual tasks, it has been<br />

also been utilized <strong>for</strong> human olfactory and gustatory tests. Indeed, a<br />

recent comparison of MLPEST with other methods to test olfactory and<br />

gustatory thresholds demonstrated reliable and precise threshold<br />

measurements (Linschoten et al. Percept. Psychophys. 63:1330, 2001).<br />

However, this comparison was limited to its application <strong>for</strong> human<br />

subjects. In order to modify this technique <strong>for</strong> olfactory testing in mice,<br />

we have adapted MLPEST methodology to the computerized<br />

olfactometer of Bodyak and Slotnick (Chem. Senses 24:637, 1999).<br />

Here we present data that demonstrate the utility of this technique in<br />

mice, and we discuss the ramifications of altering MLPEST test<br />

parameters on per<strong>for</strong>mance.<br />

This work was supported by NIH grants DC00566, DC04657 (DR)<br />

and F30 DC 5740 (AC).<br />

58 Poster [ ] Olfaction: Animal Behavior<br />

VIRUS-INDUCED BODY ODORS IN MICE<br />

Beauchamp G.K. 1, Ross S.R. 2, Curran M. 1, Hultine S. 2, Yamazaki K. 1<br />

1Monell Chemical Senses Center, Philadelphia, PA; 2Department of<br />

Microbiology, University of Pennsylvania, Philadelphia, PA<br />

Mouse Mammary Tumor Virus (MMTV) provides an attractive<br />

animal model to study the effects of infection on body odor. MMTV is<br />

a retrovirus that either is passed from mother to offspring through her<br />

milk or is transmitted genetically as an endogenous provirus.<br />

Depending on the mouse strain, multiparous infected females often<br />

develop mammary tumors. The provirus genome consists of several<br />

genes one of which, Sag codes <strong>for</strong> a superantigen that is presented by<br />

MHC Class II on B cells and acts to delete specific subsets of T cells in<br />

the host. We previously demonstrated that mice infected by MMTV by<br />

suckling on infected dams expressed a distinctive odor as shown by the<br />

ability of trained mice to discriminate odors of infected vs. non-infected<br />

controls in a Y maze. Endogenously expressed MMTV can be used to<br />

investigate mechanisms of body odor change. In the first study with<br />

endogenous virus, we found that mice transgenic <strong>for</strong> an MMTV<br />

provirus (C3H/HeN) express a different odor than control littermates.<br />

In the second study, we demonstrated that mice differing from control<br />

littermates only by expressing the Sag gene (MMTV-ORF 16) also had<br />

a distinctive body odor. This result suggests that changes in immune<br />

function, likely alterations in the T-cell repertoire, underlie body odor<br />

changes following MMTV infection. Variations in immune function<br />

following infection may thus result in specific volatile signals useful in<br />

disease diagnosis.<br />

Supported by NSF Grant#0112528<br />

15<br />

59 Poster [ ] Olfaction: Animal Behavior<br />

OLFACTORY FEAR CONDITIONING AND DISCRIMINATION<br />

IN MICE.<br />

Jones S.V. 1, Heldt S. 1, Davis M. 1, Ressler K.J. 1 1Center <strong>for</strong> Behavioral<br />

Neuroscience, Emory University, Atlanta, GA<br />

A method <strong>for</strong> producing long-lasting olfactory learning and<br />

discrimination within a single session is required <strong>for</strong> future studies of<br />

the molecular bases of olfactory memory. In previous studies, we have<br />

shown that mice can learn olfactory-cued fear as measured by both<br />

freezing and fear potentiated startle, the two most common fear<br />

measures in rodents. We first show that mice have enhanced fear<br />

potentiated startle and freezing after pairing odors with footshocks in a<br />

paradigm that supports learning, but not when the same number of<br />

odors and shock presentations are presented in an unpaired fashion.<br />

Here we also show that mice can discriminate between an odor paired<br />

with a shock versus an unpaired odor. Using 10% amyl acetate or 5%<br />

acetophenone, we examined the unconditioned startle and freezing<br />

responses of mice prior to training. During training mice received five<br />

presentations of one odor alone interspersed with five pairings of the<br />

other odor with a footshock. The following day, we tested fear<br />

potentiated startle and freezing to these odors. Acoustic startle to the<br />

untrained odor and the trained odor were non-significant be<strong>for</strong>e<br />

training. After conditioning, the fear potentiated startle response to the<br />

untrained (8%) and trained (29%) odor were significantly different<br />

(p=0.002). Additionally, freezing prior to training was non-significant,<br />

but after training, mice froze significantly more to the conditioned odor<br />

(p=0.03). Ongoing studies investigating the role of the amygdala in<br />

these behaviors will pave the way <strong>for</strong> examining the molecular<br />

mechanisms of amygdala-dependent modulation of olfactory learning.<br />

60 Poster [ ] Olfaction: Animal Behavior<br />

OPPOSING EFFECTS OF D1 AND D2 RECEPTOR<br />

ACTIVATION ON ODOR DISCRIMINATION LEARNING<br />

Mcnamara A. 1, Yue E. 1, Cleland T. 2, Linster C. 3 1Neurobiology and<br />

Behavior, Cornell University, Ithaca, NY; 2Neurobiology and Nehavior,<br />

Cornell University, Ithaca, NY; 3Cornell University*, Ithaca, NY<br />

Dopaminergic modulation of cortical activity has been implicated in<br />

the planning, initiation, and control of movements, as well as in<br />

emotional responses, motivation, and the <strong>for</strong>mation of reward<br />

associations. Additionally, there is abundant evidence <strong>for</strong><br />

dopaminergic effects on olfactory processing, and in particular <strong>for</strong> the<br />

effects of dopaminergic modulation on odor detection thresholds.<br />

Using a simultaneous olfactory discrimination task, we here show that<br />

both D1 and D2 dopamine receptors can regulate rats´ olfactory<br />

discrimination capacities, and furthermore that the effects of D1 and D2<br />

receptor activation functionally oppose one another in the per<strong>for</strong>mance<br />

of this task. Specifically, injection of either the D1 agonist SKF 38393<br />

(10 mg/kg body weight) or the D2 antagonist spiperone (0.62 mg/kg)<br />

facilitated the difficult discrimination of similar odorants but had no<br />

effect on the easier discrimination of dissimilar odorants, whereas both<br />

the D1 antagonist SCH 23390 (0.025 mg/kg) and the D2 agonist<br />

quinpirole (0.2 mg/kg) significantly impaired rats´ ability to<br />

discriminate both similar and dissimilar odorants.


61 Poster [ ] Olfaction: Animal Behavior<br />

NOT ONLY - BUT ALSO -ADRENOCEPTORS ARE INVOLVED<br />

IN EARLY ODOR PREFERENCE LEARNING IN THE RAT<br />

Mclean J.H. 1, Mccann J. 2, Darby-King A. 1, Harley C.W. 2 1Basic<br />

Medical <strong>Sciences</strong>, Memorial University of Newfoundland, St. John's,<br />

Newfoundland, Canada; 2Psychology, Memorial University of<br />

Newfoundland, St. John's, Newfoundland, Canada<br />

Early preference olfactory learning is important <strong>for</strong> the survival of<br />

infant rats because they need to associate the odor of the mother with<br />

reward (food). In the neonate, tactile stimulation activates the locus<br />

coeruleus (Nakamura et al., 1987) in the brain stem. Noradrenergic<br />

(NE) neurons in the locus coeruleus (LC) have a strong axonal<br />

projection to the olfactory bulb (Shipley et al., 1985; McLean and<br />

Shipley, 1989; McLean et al., 1991). When NE activation is paired<br />

with odor in the neonate, preference <strong>for</strong> the odor (learning) takes place<br />

(Sullivan et al., 2000;Sullivan et al., 1991;Sullivan et al., 1989;Price et<br />

al., 1998;Langdon et al., 1997). The LC β-adrenergic input has been<br />

shown to be necessary and sufficient <strong>for</strong> preference acquisition to occur<br />

in rats up to around postnatal day 10. Surprisingly, despite observations<br />

that α-adrenoceptors are present (Day et al. 1997, Winzer Serhan et al,<br />

1997) and functional (Hayar et al. 2001) in the olfactory bulb, there is<br />

relatively little known about their function in behavior. In this study, we<br />

tested the hypothesis that α-adrenoceptors have roles similar to the βadrenoceptors<br />

in early olfactory learning. In this set of experiments we<br />

showed that activation of the α1-adrenoceptors was sufficient to induce<br />

early preference learning in the rat while the α2-adrenoceptor agonists<br />

did not show an effect. The α-adrenoceptors acted in a dose-dependent<br />

manner similar to the β-adrenoceptors in early odor preference learning.<br />

This work is supported by a grant from CIHR-CEDA Regional<br />

Partnership<br />

62 Poster [ ] Olfaction: Animal Behavior<br />

THE EFFECT OF CONCENTRATION AND CONDITIONING<br />

ON ODORANT DISCRIMINATION BY THE HONEYBEE<br />

Fussnecker B.L. 1, Carlton M. 1, Wright G. 2, Smith B.H. 1 1Entomology,<br />

Ohio State University, Columbus, OH; 2Mathematical Biosciences<br />

Institute, Ohio State University, columbus, OH<br />

Naturally occurring odors used by animals <strong>for</strong> mate recognition, food<br />

identification and other purposes must be detected at concentrations that<br />

vary across several orders of magnitude. Olfactory systems must<br />

there<strong>for</strong>e have the capacity to represent odors over a large range of<br />

concentrations regardless of dramatic changes in odor salience. The<br />

stability of the representation of an odor relative to other odors across<br />

concentration has not been extensively evaluated. We tested the ability<br />

of honeybees to discriminate pure odorants across a range of<br />

concentrations at and above their detection threshold. We also<br />

examined how their ability to discriminate among odors changed as a<br />

function of the number of times they had encountered an odorant in<br />

association with an appetitive reward. Our study showed that<br />

discrimination among odorants was a function both of the concentration<br />

and the number of experiences a subject has with an odorant. We<br />

hypothesize that this arises from two separate mechanisms in the<br />

olfactory system.<br />

16<br />

63 Poster [ ] Olfaction: Animal Behavior<br />

DILTIAZEM ADMINISTERED NASALLY DECREASES FOOD<br />

INTAKE AND ATTENUATES WEIGHT GAIN IN RATS<br />

Maher T.J. 1, Amer A. 1, Adams C. 2, Chen W. 3, Weinrich K. 3<br />

1Massachusetts College of Pharmacy and Health <strong>Sciences</strong>, Boston, MA;<br />

2Compellis Pharmaceuticals, Somerville, MA; 3Longwood<br />

Pharmaceuticals, Inc., Boston, MA<br />

Energy intake is continuously influenced by many complex<br />

endogenous neurochemical systems located both centrally and<br />

peripherally, in addition to numerous external environmental stimuli,<br />

such as olfaction. As the processing of odorant in<strong>for</strong>mation by many<br />

olfactory neurons is mediated via Ca2+-currents through Ca2+<br />

channels, a novel approach at influencing the ingestive behaviors of<br />

animals might there<strong>for</strong>e involve altering olfactory acuity via Ca2+<br />

channel blockade. The present study tested the ability of a Ca2+<br />

channel blocker, diltiazem (D), to alter food intake in rats made<br />

hyperphagic. D was delivered using the intranasal (i.n.), intraperitoneal<br />

(i.p.) and oral (p.o.) routes of administration. Male Sprague Dawley rats<br />

maintained in a reversed-lighting environment, which had been fooddeprived<br />

<strong>for</strong> 4hrs at the beginning of the dark cycle, were administered<br />

different doses of D or vehicle (V) and the amount of food consumed<br />

was measured. While food intake at 1, 2 and 4 hrs post D administration<br />

was significantly decreased in a dose-dependent manner after i.n.<br />

administration, neither the i.p. nor p.o. routes significantly affected food<br />

intake. In another experiment, rats were trained to eat their daily meal<br />

during the first 4hrs at the onset of the dark cycle. Once acclimated to<br />

this schedule, daily treatment <strong>for</strong> 14 days with i.n. D or V prior to food<br />

introduction resulted in a dose-dependent attenuation of weight gain.<br />

Together these studies suggest that i.n. administration of D possesses<br />

significant anorectic activity that may have utility in influencing energy<br />

intake. Additional studies are needed to determine the exact<br />

mechanisms and sites of action of i.n. administered D.<br />

64 Poster [ ] Olfaction: Animal Behavior<br />

KV1.3-TARGETED GENE-DELETION INCREASES<br />

METABOLIC FUNCTION AND OLFACTORY ABILITY<br />

Thompson R.N. 1, Perkins R.M. 1, Parsons A.D. 2, Overton M. 2, Fadool<br />

D. 1 1Dept. Biological Sci, Prog. in Neurosci., Florida State University,<br />

Tallahassee, FL; 2Dept Nutrition, Food, & Exer Sci, Prog. in Neurosci.,<br />

Florida State University, Tallahassee, FL<br />

Mice with gene-targeted deletion (KO) of the voltage-dependent K<br />

channel, Kv1.3, have smaller glomeruli and thus were behaviorally<br />

phenotyped to understand the channel´s contribution to olfaction. Mice<br />

were monitored (8d) in environmental chambers where data are<br />

computer-acquired every 30 seconds. KO animals were found to have<br />

abnormal ingestive behaviors yet equivalent daily caloric/water intake;<br />

KOs ate more intermittently and drank larger volumes less frequently.<br />

KO animals had slightly increased metabolism and locomotor activity<br />

in the dark cycle and weighed less than aged-matched wildtype mice.<br />

KO mice were not anosmic, in fact, retrieval time to recover a hidden<br />

food item was twice as fast as that observed <strong>for</strong> wildtype mice. Odor<br />

habituation trials using complex mixtures and single alcohols indicated<br />

that KO mice can discriminate molecules differing by only 1 carbon.<br />

Food-restricted mice were trained to dig <strong>for</strong> a hidden reward paired<br />

with an odorant using a two-choice paradigm to determine odorant<br />

threshold ability. KO mice per<strong>for</strong>med the paired task more quickly and<br />

at concentrations 1,000 to 10,000 fold less than that by wildtype mice.<br />

While these tasks are influenced by memory, both genotypes per<strong>for</strong>med<br />

equivalently in object recognition testing and test of motivation <strong>for</strong><br />

object exploration. This unusual set of behaviors in the Kv1.3 KO mice<br />

suggest that Kv1.3 serves additional roles beyond shaping the resting<br />

potential. Supported by NIH DC03387(NIDCD) and NIH HL-56732.


65 Slide [ ] Trigeminal <strong>Chemoreception</strong><br />

LINGUAL TACTILE ACUITY, TASTE PERCEPTION, AND<br />

THE DENSITY AND DIAMETER OF FUNGIFORM PAPILLAE<br />

IN FEMALE SUBJECTS<br />

Chopra A. 1, Essick G. 2, Guest S. 3, Mcglone F. 4 1Cognitive<br />

Neuroscience, Unilever R&D, UK, Merseyside, United Kingdom;<br />

2University of North Carolina at Chapel Hill, Chapel Hill, NC; 3Dental<br />

Research Centre, University of North Carolina at Chapel Hill, Chapel<br />

Hill, NC; 4Cognitive Neuroscience, Unilever R&D, Merseyside, United<br />

Kingdom<br />

A growing body of evidence suggests that individuals who differ in<br />

taste perception differ in lingual tactile perception. To address this<br />

issue, spatial resolution acuity was estimated <strong>for</strong> 83 young adult<br />

females (52 Asians and 31 Caucasians) by their ability to examine with<br />

the tongue and identify embossed letters of the alphabet. Ratings of the<br />

magnitude of the bitterness of 6-n-propylthiouracil (PROP) were<br />

obtained to characterize subjects' taste perception. The density and<br />

diameter of fungi<strong>for</strong>m papillae on the anterior tongues of the Asian<br />

subjects were measured also. Subjects who rated the bitterness of PROP<br />

as very or intensely strong (supertasters) were found to be 25% more<br />

tactually acute than subjects who rated the bitterness as moderate to<br />

strong (medium tasters) and twice as acute as subjects who rated it as<br />

barely detectable or weak (non-tasters; P


69 Slide [ ] Trigeminal <strong>Chemoreception</strong><br />

FATTY ACIDS INHIBIT DELAYED RECTIFYING K<br />

CHANNELS IN ISOLATED TRIGEMINAL NEURONS.<br />

Gilbertson T.A. 1, Klein J.T. 1, Farmer-George M. 1, Hansen D.R. 1, Simon<br />

S.A. 2 1Biology & Center <strong>for</strong> Integrated BioSystems, Utah State<br />

University, Logan, UT; 2Anesthesiology, Duke University, Durham, NC<br />

Chemosensory cues <strong>for</strong> fat have been shown to be mediated by fatty<br />

acids (FA) acting directly on subtypes of delayed rectifying K (DRK)<br />

channels in taste receptor cells (TRCs). Since it is generally believed<br />

that texture is important <strong>for</strong> oral fat recognition and that texture<br />

perception is mediated, at least in part, by trigeminal (TG) innervation,<br />

we have examined the effects of fatty acids (0.1-10 µM) on isolated rat<br />

TG neurons using whole-cell patch clamp recording. TG ganglia were<br />

removed and individual TG neurons were isolated by enzymatic<br />

methods and placed into culture <strong>for</strong> 24-72 h. Patch recordings were<br />

made be<strong>for</strong>e, during and after application of a variety of fatty acids<br />

including saturated (SFA), monounsaturated (MUFA) and<br />

polyunsaturated (PUFA) <strong>for</strong>ms. Similar to our results in TRCs, PUFAs<br />

caused a reversible, time-dependent inhibition of DRK channels in TG<br />

neurons, consistent with an open-channel block leading to cell<br />

activation. In contrast to fungi<strong>for</strong>m TRCs, which respond specifically to<br />

PUFAs, TG neurons were much less specific and responded to a variety<br />

of fatty acid types, including some MUFA and SFA, in a similar<br />

fashion. Thus, FA effects on DRK channels may not only mediate the<br />

taste of fat, but may also contribute to the perception of its textural<br />

properties via activation of oral TG fibers. Currently, we are using<br />

quantitative PCR to compare the expression of DRK channels in TG<br />

neurons with TRCs to determine the source of the FA specificity<br />

differences. Supported by NIH DK59611(TAG), DC01065 (SAS).<br />

70 Symposium [ ] Receptors: Choosing Genes, Targeting<br />

Axons, Detecting Chemicals<br />

PERCEPTION OF CHEMICAL CUES AND NAVIGATION IN C.<br />

ELEGANS<br />

Bargmann C.I. 1 1Anatomy, University of Cali<strong>for</strong>nia, San Francisco,<br />

San Francisco, CA<br />

Behavior arises from the interplay between the environment and<br />

intrinsic properties of neurons and neural circuits. To understand how<br />

the genetics and development of the nervous system contribute to<br />

specific behaviors, we are studying olfactory system in the nematode C.<br />

elegans. C. elegans senses hundreds of different compounds,<br />

discriminates between them, and generates different behaviors in<br />

response to different odors. It is possible to define the specific neurons<br />

that generate these behaviors, since the C. elegans nervous system<br />

consists of just 302 neurons that have reproducible functions,<br />

morphologies and synaptic connections. Previous studies have<br />

generated an understanding of the methods by which animals detect and<br />

respond to a single sensory stimulus. In C. elegans, odors are detected<br />

by over 1000 G protein-coupled odorant receptors. Individual olfactory<br />

neurons express multiple receptor genes, allowing a few cells to detect<br />

many odors. A given sensory neuron is primarily dedicated to a single<br />

behavioral task, such as attraction or repulsion. We are now asking how<br />

animals navigate through complex sensory environments using multiple<br />

odors or sensory inputs. For these studies, we have focused on complex<br />

natural stimuli that should be present in the soil environment, such as<br />

different bacterial foods, natural physical stimuli, and other animals<br />

(social groups). C. elegans shows unexpected sophistication in its<br />

behavior when it faces ecologically relevant challenges like pathogenic<br />

bacteria or metabolic stress. Using the wiring diagram, we are<br />

identifying the circuits <strong>for</strong> navigation behavior and asking how sensory<br />

inputs regulate those circuits.<br />

18<br />

71 Symposium [ ] Receptors: Choosing Genes, Targeting<br />

Axons, Detecting Chemicals<br />

THE BIOLOGY OF SWEET, BITTER AND UMAMI TASTE<br />

Zuker C.S. 1 1Section of Neurobiology, University of Cali<strong>for</strong>nia, San<br />

Diego, La Jolla, CA<br />

72 Symposium [ ] Receptors: Choosing Genes, Targeting<br />

Axons, Detecting Chemicals<br />

INTERNAL REPRESENTATIONS OF THE OLFACTORY<br />

WORLD<br />

Wang J. 1, Wong A.M. 2, Axel R. 3 1Neurobiology and Behavior,<br />

Columbia University, New York, NY; 2Department of Biochemistry and<br />

Molecular Biophysics, Columbia University, New YOrk, NY;<br />

3Biochemistry and Molecular Biophysics, Columbia University, New<br />

York, NY<br />

Olfactory perception requires the recognition of a vast repertoire of<br />

odorants in the periphery and central neural mechanisms that allow the<br />

discrimination of odors. The organization of the peripheral olfactory<br />

system appears remarkably similar in fruit flies and mammals. The<br />

convergence of like axons into discrete glomerular structures provides<br />

an anatomic map in the antennal lobe. How does the anatomic map<br />

translate into a functional map? We have developed a sensitive imaging<br />

system in the Drosophila brain that couples two-photon microscopy<br />

with the specific expression of the calcium-sensitive fluorescent<br />

protein, G-CaMP, to examine neural activity. At natural odor<br />

concentrations, each odor elicits a distinct and sparse pattern of activity<br />

that is conserved in different flies. We have combined Ca2+ imaging<br />

with electrical recordings to demonstrate the faithful propagation of the<br />

glomerular map by projection neurons that innervate the protocerebrum.<br />

The quality of an odor may there<strong>for</strong>e be reflected by defined spatial<br />

patterns of activity, first in the antennal lobe and ultimately in higher<br />

olfactory centers. We have identified a spatially invariant sensory map<br />

in the fly protocerebrum that is divergent and no longer exhibits the<br />

insular segregation of like axons observed in the antennal lobe. This<br />

organization provides the opportunity <strong>for</strong> the integration of multiple<br />

glomerular inputs by hierarchical cell assemblies in the protocerebrum.


73 Poster [ ] Salt and Sour Taste<br />

EXPRESSION OF THE AMILORIDE-SENSITIVE EPITHELIAL<br />

SODIUM CHANNEL IN THE MOUSE TASTE PAPILLAE<br />

Shigemura N. 1, Sadamitsu C. 2, Yasumatsu K. 1, Yoshida R. 1, Ninomiya<br />

Y. 1 1Section of Oral Neuroscience, Graduate school of Dental Science,<br />

Kyushu University, Fukuoka, Japan; 2Graduate School of<br />

Pharmaceutical <strong>Sciences</strong>, University of Tokyo, Tokyo, Japan<br />

It is proposed that amiliride-sensitive epithelial Na + Channels<br />

(ENaCs) are involved in taste signal transduction. Electrophysiological<br />

studies in C57 BL mice demonstrated that responses to NaCl are<br />

inhibited by amilolide in the choda tympani (CT) nerve but not in the<br />

glossopharyngeal (IXth) nerve, suggesting lack of amiloride sensitivity<br />

(AS) in the posterior tongue region innervated by IXth nerve. The AS<br />

also differs among inbred mouse strains. Judging from amiloride<br />

inhibition of NaCl responses of the CT nerve, the BALB strain showed<br />

very weak AS even in the anterior tongue. In this study, by using in situ<br />

hybridization (ISH) and semiquantitative RT-PCR techniques, we<br />

examined expression of three subunits of ENaC (alpha, beta, gamma) in<br />

the fungi<strong>for</strong>m papillae (FP), circumvallate papilla (CP) and tongue<br />

epithelial tissue without taste papillae (ET) in C57BL and BALB mice.<br />

The results demonstrated that signals <strong>for</strong> alpha subunit of ENaC were<br />

clearly detected in some spindle-shape cells in both FP and CP, whereas<br />

those <strong>for</strong> beta and gamma subunits were clearly detected in some taste<br />

cells in FP, but only slightly in CP. In the ET, all three subunits were<br />

detected. These results together with those reported by previous studies<br />

suggest that expression patterns <strong>for</strong> the three subunits of ENaC may<br />

contribute to the tongue regional difference in AS in mice. With regard<br />

to the strain difference in AS, we are examining the gene polymorphism<br />

<strong>for</strong> each subunits of ENaC between C57BL and BALB strains by<br />

sequencing analysis.<br />

74 Poster [ ] Salt and Sour Taste<br />

QUANTITATIVE PCR ANALYSIS OF THE ALDOSTERONE-<br />

REGULATED SALT TRANSDUCTION PATHWAY IN TASTE<br />

CELLS.<br />

Burks C.A. 1, Hansen D.R. 1, Gilbertson T.A. 1 1Biology & Center <strong>for</strong><br />

Integrated BioSystems, Utah State University, Logan, UT<br />

Evidence suggests that sodium salt taste transduction pathways<br />

involving the epithelial sodium channel (ENaC) are responsive to<br />

aldosterone (ALDO) via its well-documented late genomic effects (12-<br />

48 h). Using RT-PCR, we previously identified a number of early (


77 Poster [ ] Salt and Sour Taste<br />

AMILORIDE DISRUPTS NACL VS. KCL TASTE<br />

DISCRIMINATION IN INBRED MICE WHETHER THEIR<br />

CHORDA TYMPANI NERVES ARE AMILORIDE SENSITIVE<br />

OR NOT<br />

Eylam S. 1, Spector A.C. 1 1Department of Psychology and Center <strong>for</strong><br />

Smell and Taste, University of Florida, Gainesville, FL<br />

In rodent taste bud cells, NaCl is transduced via one pathway<br />

selective <strong>for</strong> Na + and disrupted by the lingual application of the<br />

epithelial Na + channel blocker amiloride, and another less cationselective<br />

and amiloride-insensitive pathway(s). In rats, amiloride<br />

appears to render NaCl qualitatively indistinguishable from KCl in a<br />

dose-dependent fashion. In mice, amiloride has been shown to reduce<br />

chorda tympani (CT) responses to NaCl in some strains (e.g. C57BL/6J<br />

(B6)) but not others (e.g. 129, DBA/2 (D2) and BALB/c (BALB)). We<br />

used a two-response operant task to train mice from these 4 strains<br />

(7/strain) in a gustometer to discriminate between 5-lick samples of<br />

NaCl and KCl with concentration (0.1-1 M) varied to render intensity<br />

irrelevant. Unexpectedly, the overall per<strong>for</strong>mance of the B6 mice was<br />

significantly lower than that of the other strains. All mice were then<br />

tested with the stimuli adulterated with a descending order of amiloride<br />

concentrations (0.1-100 µM) with control sessions interposed. In all 4<br />

strains, overall per<strong>for</strong>mance dropped to chance at 100 µM amiloride<br />

and sigmoidally improved as the adulterating amiloride concentration<br />

was lowered. Per<strong>for</strong>mance on NaCl trials was much more affected than<br />

that on KCl trials. Thus, the ability to discriminate between NaCl and<br />

KCl depends on an amiloride-sensitive transduction pathway in the 4<br />

strains tested here. Possibly, D2, 129, and BALB (and perhaps B6)<br />

mice have amiloride-sensitive taste receptor cells innervated by<br />

gustatory nerves other than the CT. Supported by NIDCD R01-<br />

DC04574.<br />

78 Poster [ ] Salt and Sour Taste<br />

SUPRATHRESHOLD INTENSITY DISCRIMINATION OF<br />

NACL IN RATS WITH CHORDA TYMPANI OR<br />

GLOSSOPHARYNGEAL NERVE TRANSECTION<br />

Colbert C.L. 1, Garcea M. 1, Spector A.C. 1 1Psychology Department and<br />

Center <strong>for</strong> Smell and Taste, University of Florida, Gainesville, FL<br />

In rats, chorda tympani nerve (CT) transection (CTX) but not<br />

glossopharyngeal nerve (GL) transection (GLX) raises NaCl taste<br />

detection threshold, but the effect of CTX or GLX on suprathreshold<br />

intensity discrimination is unknown. We examined whether CTX or<br />

GLX in male Sprague-Dawley rats would disrupt a presurgically trained<br />

NaCl intensity discrimination in a two-lever operant task. Thirsty rats<br />

were required to correctly press one lever in response to a standard<br />

concentration and the other lever in response to higher comparison<br />

concentrations <strong>for</strong> water reward. The hit rate <strong>for</strong> comparison stimuli<br />

was adjusted <strong>for</strong> the standard false alarm rate and a logistic function<br />

was fit to the concentration-per<strong>for</strong>mance data. The concentrations at the<br />

curve midpoints (c) were used to compare differential sensitivity across<br />

groups. The mean c <strong>for</strong> the CTX group (n=4) significantly differed<br />

from that <strong>for</strong> the sham-operated (SHAM; n=4) and GLX (n=3) groups<br />

postsurgically with a 0.05 M NaCl standard, though the magnitude of<br />

difference was only 0.16 log units. The corresponding Weber<br />

10<br />

Fractions (WF) were significantly higher in the CTX group (1.33)<br />

compared with the SHAM (.64) and GLX (.64) groups. In contrast, the<br />

cs nor WFs (CTX=.70; GLX=.67; SHAM=.49) significantly differed<br />

across groups when the standard was 0.1 M NaCl. These findings add<br />

to a growing list of taste-related behavior <strong>for</strong> which the GL is<br />

unnecessary, but the small effect of CTX on NaCl intensity<br />

discrimination is in contrast with its more striking effects on absolute<br />

detection. Supported by NIH R01-DC01628.<br />

20<br />

79 Poster [ ] Salt and Sour Taste<br />

RELATIVE EFFECTS OF TRANSECTION OF THE<br />

GUSTATORY BRANCHES OF THE 7TH AND 9TH CRANIAL<br />

NERVES ON NACL DETECTABILITY IN RATS<br />

Blonde G.D. 1, Garcea M. 1, Spector A.C. 1 1Department of Psychology<br />

and Center <strong>for</strong> Smell and Taste, University of Florida, Gainesville, FL<br />

Chorda tympani nerve (CT) transection in rats severely impairs NaCl<br />

taste detection. Higher concentrations of NaCl are nonetheless<br />

detectable suggesting that the remaining gustatory nerves can maintain<br />

some degree of salt sensibility. We used a two-response operant taste<br />

detection task to presurgically train male Sprague-Dawley rats and<br />

measure NaCl sensitivity in a gustometer. A modified descending<br />

method of limits procedure was used in which a single concentration,<br />

pitted against water in a session, was progressively lowered across test<br />

days. Logistic functions based on overall percentage correct were<br />

derived. The concentrations at the curve midpoints (c) were used to<br />

compare NaCl sensitivity be<strong>for</strong>e and after sham surgery (SHAM, n=5),<br />

transection of both the CT and greater superficial petrosal nerve (GSP)<br />

(7X; n=6), or transection of the glossopharyngeal nerve (9th ) (9X, n=4).<br />

The value of c did not significantly change after surgery <strong>for</strong> the SHAM<br />

and 9X groups. In contrast, the 7X surgery significantly raised c by<br />

~2.50 log units, a shift greater than that reported <strong>for</strong> CT transection.<br />

10<br />

These rats were still able to detect very high concentrations suggesting<br />

that the 9th or superior laryngeal nerve might support NaCl detection at<br />

these stimulus intensities, although the role of trigeminal cues cannot be<br />

dismissed. Histology is currently underway. These results suggest that<br />

the GSP contributes to NaCl sensitivity in the rat and confirm the<br />

primacy of the 7th nerve relative to the 9th nerve in sensibility to NaCl.<br />

Supported by NIH R01-DC01628.<br />

80 Poster [ ] Salt and Sour Taste<br />

ASIC2 IS NOT NECESSARY FOR SOUR TASTE IN MICE<br />

Richter T.A. 1, Dvoriantchikov G. 1, Chaudhari N. 2, Roper S.D. 2<br />

1Physiology & Biophysics, University of Miami School of Medicine,<br />

Miami, FL; 2Physiology & Biophysics and Program in Neuroscience,<br />

University of Miami School of Medicine, Miami, FL<br />

The acid-sensitive cation channel, ASIC2, is widely believed to be a<br />

receptor <strong>for</strong> acid (sour) taste in mammals, based on its physiological<br />

properties and expression in rat taste cells. Thus, we evaluated acid<br />

taste responses in wild type and ASIC2 knockout mice. Ca 2+ imaging<br />

experiments were carried out using confocal microscopy on slices of<br />

circumvallate papillae to which taste stimuli were applied focally at the<br />

taste pore. Surprisingly, taste cells from ASIC2 knockout mice<br />

exhibited normal physiological responses to 20-100mM citric acid. The<br />

incidence (number of acid-responsive cells) and response amplitude<br />

(increase in [Ca 2+ ]i) were identical across the two genotypes. Hence, we<br />

explored the expression pattern of the four channels, ASIC1-4, in taste<br />

buds from both rat and mouse. Using RT-PCR, we detected expression<br />

of ASIC1 and ASIC3, but not ASIC4, in mouse and rat taste buds and<br />

non-sensory lingual epithelium. Interestingly, we found that mRNA <strong>for</strong><br />

ASIC2 is not expressed at significant levels in mouse taste buds,<br />

although we observed its robust expression in rat taste buds. Our results<br />

indicate that ASIC2 is not required <strong>for</strong> acid taste in mice. The findings<br />

also highlight that there may be important differences in taste<br />

transduction mechanisms between mice and rats. Supported by<br />

NIH/NIDCD grants 2RO1 DC00374 (SDR) and 1R21 DC5500 (NC).


81 Poster [ ] Salt and Sour Taste<br />

COLD INHIBITS SOUR TASTE TRANSDUCTION<br />

Defazio R.A. 1, Richter T.A. 1, Roper S.D. 1 1Physiology and Biophysics,<br />

University of Miami, Miami, FL<br />

The taste transduction mechanism underlying sour taste remains to be<br />

elucidated. The basic mechanism is believed to be an acid-sensitive ion<br />

channel, the activation of which depolarizes the membrane potential.<br />

Membrane depolarization activates voltage-gated calcium channels<br />

(VGCCs) and allows Ca2+ influx, thereby increasing intracellular<br />

calcium ([Ca2+ ] ). Increased [Ca i 2+ ] presumably initiates<br />

i<br />

neurotransmitter release. Many ion channels, especially acid-gated<br />

channels, are sensitive to changes in temperature. Thus, we reasoned<br />

that the temperature sensitivity of sour taste transduction might provide<br />

clues to the identity of transduction channels in acid-responsive taste<br />

cells. We examined the effect of temperature on tastant-induced<br />

changes in [Ca2+ ] in mouse vallate taste cells using confocal<br />

i<br />

fluorescence microscopy and acutely prepared lingual slices. At 30°C,<br />

bath application of 100 mM KCl depolarized cells and activated<br />

VGCCs, resulting in increased [Ca2+ ] in a subpopulation of taste cells.<br />

i<br />

In a subset of these KCl-responsive cells, focal application of 100 mM<br />

citric acid produced robust increases in [Ca2+ ] . When the bath<br />

i<br />

temperature was lowered to 17°C, the response to citric acid was<br />

reduced 97±2% (n=4), that is, effectively abolished. In contrast, the<br />

KCl response was reduced by less than half (43±4%, n=5). Our results<br />

suggest that cold inhibits sour taste transduction at the initial<br />

transduction channel rather than by blocking downstream Ca2+ influx<br />

via VGCCs. This observation may be diagnostic of certain proton-gated<br />

channels. Supported by grant DC00374 (SDR).<br />

82 Poster [ ] Salt and Sour Taste<br />

NEURAL ADAPTATION TO ACID STIMULI IS MODULATED<br />

BY THE BASOLATERAL NA+-H+ EXCHANGER-1 (NHE-1) IN<br />

FUNGIFORM TASTE RECEPTOR CELLS (TRCS)<br />

Desimone J.A. 1, Lyall V. 1, Phan T.T. 1, Vinnikova A.K. 2, Heck G.L. 1<br />

1Physiology, Virginia Commonwealth University, Richmond, VA;<br />

2Internal Medicine, Virginia Commonwealth University, Richmond, VA<br />

The role of NHE-1 in neural adaptation to acidic stimuli was<br />

investigated by intracellular pH (pH ) measurements in polarized rat<br />

i<br />

fungi<strong>for</strong>m TRCs and chorda tympani (CT) nerve recordings. In<br />

- + CO /HCO -free media (pH 7.4), basolateral Na removal decreased<br />

2 3<br />

TRC pH and zoniporide, a specific NHE-1 blocker, inhibited the Na i + -<br />

induced decrease in pH . The TRC pH recovery rate from NH Cl pulses<br />

i i 4<br />

was inhibited by basolateral zoniporide with a K of 0.33 µM.<br />

i<br />

Basolateral ionomycin reversibly increased TRC Ca2+ , resting pH , and i<br />

the pH recovery rate from an NH Cl pulse. These effects of Ca i 4 2+ were<br />

blocked by zoniporide. The lingual application of zoniporide increased<br />

the magnitude of the CT responses to acetic acid and CO , but not to<br />

2<br />

HCl. Lingual application of ionomycin did not affect the phasic part of<br />

the CT responses to acidic stimuli, but decreased the tonic part by 50%<br />

of control over a period of about 1 min. This increased adaptation in the<br />

CT response was inhibited by zoniporide. Lingual application of 8-<br />

CPT-cAMP increased CT responses to HCl, but not to CO , and acetic<br />

2<br />

acid. In the presence of cAMP, ionomycin increased sensory adaptation<br />

to HCl, CO , and acetic acid. Thus, cAMP and Ca 2 2+ independently<br />

modulate CT responses to acidic stimuli. While cAMP enhances TRC<br />

apical H + entry and CT responses to strong acid, an increase in Ca2+ activates NHE-1, and increases neural adaptation to all acidic stimuli.<br />

Supported by NIDCD Grants DC-02422 and DC-00122.<br />

21<br />

83 Poster [ ] Salt and Sour Taste<br />

BDNF HAPLOINSUFFICIENT MICE HAVE SELECTIVE<br />

TASTE LOSS FOR SOUR<br />

Barrows J. 1, Kinnamon S.C. 2, Vigers A. 1, Finger T.E. 1 1Cell and<br />

Developmental Biology, University of Colorado Health <strong>Sciences</strong><br />

Center, Denver, CO; 2Colorado State University, Fort Collins, CO<br />

Taste receptor cells are classified according to morphological and<br />

histochemical criteria. Type II cells express components of the<br />

phospholipase C signaling pathway (Clapp et al. 2004), required <strong>for</strong><br />

bitter, sweet, and umami transduction (Zhang et al., 2002). Type III<br />

cells express brain-derived neurotrophic factor (BDNF) and voltagegated<br />

calcium channels (Medler et al., 2003). A recent study suggested<br />

that voltage-gated calcium channels may be required <strong>for</strong> sour<br />

transduction (Richter et al., 2003). Taken together, these findings<br />

suggest that Type III cells may mediate sour transduction. The current<br />

study tested whether mice haploinsufficient <strong>for</strong> BDNF have reduced<br />

sensitivity to sour tastants. The haploinsufficient mice utilized in these<br />

studies have a portion of the coding region of BDNF replaced by lacZ.<br />

Taste buds of these BDNF/lacZ mice are smaller than wild type<br />

littermates (Yee et al., 2003). Mice were tested using 24-hour twobottle<br />

taste preference <strong>for</strong> the response to citric acid (1 mM to 30 mM),<br />

the artifical sweetener, SC45647 (3-100 uM), and the bitter substance,<br />

denatonium benzoate (0.01 mM to 3 mM). The BDNF/lacZ mice<br />

appear to be one-half log unit less sensitive to citric acid compared to<br />

wild type animals but do not differ from wild types in sensitivity to<br />

denatonium or SC45647. These data suggest that sour taste may be<br />

specifically impacted in BDNF haploinsufficient mice and suggest that<br />

BDNF-expressing Type III taste cells are necessary <strong>for</strong> sour taste<br />

transduction. Funded by NIH Grants DC006070, DC00244 and P30<br />

DC04657.<br />

84 Poster [ ] Gustatory Processing<br />

THE AMILORIDE-SENSITIVE NA-CHANNEL AND NACL-<br />

QUININE MIXTURE INHIBITION IN THE CHORDA<br />

TYMPANI<br />

Formaker B.K. 1, Dowling R.S. 1, Frank M.E. 2 1Oral Diagnosis,<br />

University of Connecticut, Farmington, CT; 2BioStructure & Function,<br />

University of Connecticut, Farmington, CT<br />

NaCl inhibits chorda tympani (CT) taste responses to quinine·HCl<br />

(QHCl) when the two compounds are presented together in a mixture<br />

(Formaker et al., 1996). We tested the hypothesis that activation of the<br />

amiloride-sensitive Na + channel plays a role in the inhibition of QHCl<br />

responses by treating the tongue with amiloride and by substituting K +<br />

<strong>for</strong> Na + . We recorded taste responses from the CT of 6 golden hamsters<br />

(Mesocricetus auratus) be<strong>for</strong>e and after lingual treatment with 30 µM<br />

amiloride. Taste stimuli were 30 mM NaCl, 30 mM QHCl, 50 mM KCl<br />

and the binary combinations of NaCl or KCl with QHCl. Responses to<br />

500 mM NH 4 Cl were used to normalize response measurements.<br />

Amiloride treatment effectively eliminated differences between KCl,<br />

NaCl, and their respective mixtures with QHCl. Without amiloride,<br />

responses to NaCl-QHCl mixtures equaled responses to NaCl alone.<br />

With amiloride, the response to the NaCl-QHCl mixture was greater<br />

than the response to either mixture component and less than the<br />

response predicted by additivity, F(3,15) = 46.1, p


85 Poster [ ] Gustatory Processing<br />

CHEMICAL AND THERMAL RESPONSES OF GUSTATORY<br />

NEURONS IN THE GENICULATE GANGLION<br />

Breza J.M. 1, Curtis K. 1, Contreras R.J. 1 1Department of Psychology and<br />

Program in Neuroscience, Florida State University, Tallahassee, FL<br />

We examined chemical and thermal responses of gustatory neurons<br />

(N=18) in the geniculate ganglion using extracellular single-cell<br />

electrophysiology. Cells were categorized by responses to lingual<br />

stimulation with the 4 classic tastants: NaCl (100mM), QHCl (20mM),<br />

sucrose (500mM), and citric acid (10mM) at a baseline temperature of<br />

25°C. Responses to temperature change of 1°C/s/15s (±15°C) also<br />

were measured. Finally, responses to the 4 tastants were evaluated after<br />

adaptation to 40°C and to 10°C. We recorded from 1 Na-specialist<br />

(5.6%), 1 QHCl-best (5.6%), 2 acid-best (11.1%) and 14 Na-generalists<br />

(77.8%). Temperature change affected firing in only 3/18 cells (16.7%)<br />

and only when cooled from 40° to 25°C. Specifically, 2 Na-generalists<br />

increased firing by 1.1spikes/s (sps) during cooling and the QHCl-best<br />

cell increased by 2.9sps. After 3-5min adaptation to 10°C, responses to<br />

all chemical stimuli decreased in all cells, with the responses to the best<br />

stimulus reduced by 2.5±0.3sps. The most robust decrease in activity<br />

occurred in the QHCl-best cell (-4.9sps); acid best cells decreased firing<br />

by -3.9sps. In contrast, after 3-5min adaptation to 40°C, responses<br />

increased to the best stimulus in 10/16 cells tested (62.5%): 2/2 acidbest<br />

(+1.9sps) and 8/13 Na-generalists (+2.0±0.4sps). Responses<br />

decreased or were not affected in the remaining 6 cells: 5/13 Nageneralists<br />

(-0.8±0.5sps) and the QHCl-best cell (-1.5sps). Thus,<br />

temperature differentially affects responses of geniculate ganglion<br />

neurons to taste stimuli. Supported by NIH Grant DC 04875.<br />

86 Poster [ ] Gustatory Processing<br />

FEEDING RELATED PEPTIDES: INVOLVEMENT WITH THE<br />

GUSTATORY SYSTEM IN THE BRAIN OF GOLDFISH<br />

(CARASSIUS AURATUS)<br />

Huesa G. 1, Finger T. 2 1University of Colorado Health <strong>Sciences</strong> Center,<br />

Denver, CO; 2Cell and Developmental Biology, University of Colorado<br />

Health <strong>Sciences</strong> Center, Denver, CO<br />

Melanin Concentrating Hormone (MCH) and Hypocretins (Hcrt) -<br />

also called orexins-, are peptidergic neurotransmitters expressed<br />

primarily in neurons located in the lateral hypothalamus. This region is<br />

known to have a critical role in control of feeding and these peptides<br />

actively participate in the modulation (stimulation) of food intake. To<br />

test whether these orexigenic peptidergic systems are involved in the<br />

gustatory in<strong>for</strong>mation processing we examined their distribution in<br />

gustatory nuclei of the CNS of goldfish. For this purpose we employed<br />

immunocytochemical techniques using polyclonal antisera directed<br />

against MCH or Hcrt2. The reaction was developed with<br />

diaminobenzidine to enhance the intensity of the labeling. The<br />

distribution of these two peptides was similar to those described in<br />

mammals, with cell bodies in the hypothalamic region and fibers<br />

projecting rostrally and caudally. Hcrt2 immunoreactive fibers are<br />

located mainly in preoptic areas, hypothalamus, a few mesencephalic<br />

regions and the dorsal horn in the spinal cord, but are scarce in taste<br />

related brainstem nuclei. In contrast, MCH fibers were present in many<br />

encephalic regions including the vagal gustatory lobe. These MCH<br />

fibers occur mainly in layer IV of the sensory zone of the vagal lobe,<br />

where many primary gustatory afferents terminate. In addition,<br />

occasional MCH fibers could be observed crossing the sensory zone, in<br />

layer IX and also in fiber and motor layers. These results support the<br />

concept that MCH but not Hypocretins may act as a central modulator<br />

of gustatory processing in goldfish. Supported by NIDCD Grant:<br />

DC00147<br />

22<br />

87 Poster [ ] Gustatory Processing<br />

EXPRESSION OF NEUROTRANSMITTER RECEPTORS<br />

WITHIN THE NUCLEUS OF THE SOLITARY TRACT OF THE<br />

HAMSTER<br />

Ye M.K. 1, Rubrum A.M. 1, Christy R.C. 1, Smith D.V. 1 1Anatomy &<br />

Neurobiology, University of Tennessee Health Science Center,<br />

Memphis, TN<br />

The nucleus of the solitary tract (NST) processes taste in<strong>for</strong>mation<br />

from the periphery that is modulated by descending projections from<br />

several areas of the <strong>for</strong>ebrain. Glutamate receptors (GluR) mediate<br />

peripheral excitation of all taste-responsive NST neurons: Both<br />

AMPA/kainate and NMDA antagonists block driven activity. Subsets<br />

of these neurons are excited by microinjection of substance P and<br />

inhibited by local infusion of γ-aminobutyric acid (GABA) or metenkephalin.<br />

To further delineate the neural circuitry involved in these<br />

modulatory effects, we are examining the expression of a number of<br />

neurotransmitter receptors in the NST. Serial coronal or horizontal 50-<br />

µm sections through the hamster brainstem were processed <strong>for</strong><br />

immunoreactivity to antibodies directed against µ- and δ-opioid<br />

receptors, glutamate receptor subtypes GluR-1, -3 and -4 (AMPA) and<br />

GluR-6 (kainate), the tachykinin receptor NK1, and the GABAA receptor. A given brain was processed using the ABC method <strong>for</strong> one<br />

of these eight antibodies. Cells within the gustatory region of the NST<br />

expressed GluR-1, -3, -4 or –6, the NK1, or the GABA receptor.<br />

A<br />

Afferent fibers of the VIIth and IXth nerves entering the solitary tract<br />

expressed the µ-opioid receptor, whereas the δ-opioid receptor was seen<br />

in cells of the NST, including the gustatory zone. This pattern of<br />

expression would suggest a substrate <strong>for</strong> both pre- and postsynaptic<br />

inhibition of gustatory neurons. Further work will examine receptor<br />

expression in physiologically characterized NST neurons. Supported<br />

by NIDCD DC000066 to DVS.<br />

88 Poster [ ] Gustatory Processing<br />

PAIRED PULSE FACILITATION AND DEPRESSION<br />

OBSERVED IN TASTE CELLS IN THE RAT NUCLEUS OF<br />

THE SOLITARY TRACT FOLLOWING<br />

GLOSSOPHARYNGEAL NERVE STIMULATION<br />

Hallock R. 1, Di Lorenzo P. 1 1Psychology, State University of New York<br />

at Binghamton, Binghamton, NY<br />

Evoked responses from paired pulse stimulation of the<br />

glossopharyngeal (GP) nerve was studied in taste responsive cells in the<br />

nucleus of the solitary tract (NTS) of anesthetized rats. Adult male rats<br />

were anesthetized with urethane (1.4 g/kg) and Nembutal (25 mg/kg)<br />

and prepared <strong>for</strong> unilateral electrical stimulation of the lingual branch<br />

of the GP nerve and <strong>for</strong> electrophysiological recording from the<br />

ipsilateral NTS. Once a taste-responsive cell was isolated, responses to<br />

taste stimuli (0.1 M NaCl, 0.5 M sucrose, 0.01 M quinine HCl, 0.01 M<br />

HCl) and electrical stimulation of the GP nerve were recorded. Up to<br />

100 Pairs of pulses were delivered to the GP nerve at various interpulse<br />

intervals and responses from cells were recorded. For each cell, the<br />

threshold <strong>for</strong> stimulation of the nerve that evoked a response was<br />

determined and a slightly higher current level was used <strong>for</strong> subsequent<br />

stimulation. Electrical pulses were .2 ms in duration and the maximum<br />

current level was 1.5 mA. Thus far, both paired pulse facilitation and<br />

depression have been demonstrated in taste cells in the NTS in response<br />

to stimulation of the GP. Much like paired pulse depression that has<br />

been previously demonstrated with stimulation of the chorda tympani<br />

nerve, paired pulse facilitation and depression in the GP may be<br />

mechanisms by which cells integrate incoming in<strong>for</strong>mation, modify the<br />

responses of other cells, or enable the system to increase contrast<br />

among taste stimuli.<br />

Supported by NIDCD grant 1-RO1-DC005219 to PMD.


89 Poster [ ] Gustatory Processing<br />

SINGLE TASTE-RESPONSIVE NEURONS IN THE NUCLEUS<br />

OF SOLITARY TRACT PROJECT AXONS TO BOTH<br />

PARABRACHIAL AND HYPOGLOSSAL NUCLEI: AN IN-<br />

VIVO INTRACELLULAR RECORDING AND LABELING<br />

STUDY<br />

Li C.X. 1, Li C.S. 1, Smith D.V. 1, Waters R.S. 1 1Anatomy and<br />

Neurobiology, University of Tennessee Health Science Center,<br />

Memphis, TN<br />

We examined the axonal projections of individual taste-responsive<br />

neurons of the nucleus of the solitary tract (NST) of hamsters using<br />

intracellular recording and labeling. The activity of NST neurons was<br />

recorded intracellularly to determine their intrinsic firing patterns and<br />

their responses to lingual stimulation with anodal current and with<br />

sucrose, NaCl, citric acid, and quinine. Recorded cells were labeled<br />

with 2% biocytin (500 ms, 1 na, 1.4 Hz, 6–40 min); a minimum of 2.5<br />

hrs was imposed between cell labeling and perfusion to permit axonal<br />

transport. Hamsters were given lethal injections of Nembutal and<br />

perfused with chilled 4% para<strong>for</strong>maldehyde. Brains were removed,<br />

refrigerated in fixative overnight, and sectioned at 100 µm in a<br />

horizontal or sagittal plane. Sections were incubated in ABC and<br />

reacted with DAB to reveal the labeled cell and counterstained with<br />

cytochrome oxidase to visualize brainstem nuclei and tracts. To date,<br />

we have injected six cells (3 multipolar, 2 elongate, 1 bipolar-like) and<br />

traced the axonal projections to medial and lateral divisions of the<br />

parabrachial nuclei (PbN) <strong>for</strong> three cells. In two cells, NST axons<br />

projected to both the PbN and the hypoglossal nucleus. NST axons<br />

contained bouton-like swellings along their axonal pathway as well as<br />

in their terminal fields. These results suggest that a single class of NST<br />

cells can provide output to three or more separate synaptic targets.<br />

Supported by NSF IBN-9724092 to RSW and NIH DC000066 to<br />

DVS.<br />

90 Poster [ ] Gustatory Processing<br />

TASTE-EVOKED RESPONSES IN THE NUCLEUS OF THE<br />

SOLITARY TRACT OF C57BL/6BYJ AND 129P3/J MICE.<br />

Mccaughey S. 1, Bachmanov A. 1 1Monell Chemical Senses Center,<br />

Philadelphia, PA<br />

Although mice offer several advantages <strong>for</strong> examining the role of<br />

genetic factors in taste sensation, gustatory responses have been<br />

recorded only from the periphery in this species. In the present<br />

experiment, the properties of cells in the first central gustatory relay, the<br />

nucleus of the solitary tract (NST), were determined in male<br />

C57BL/6ByJ (B6) and 129P3/J (129) mice. Animals were anesthetized,<br />

a surgery was per<strong>for</strong>med to expose the surface of the brainstem, and<br />

neural activity was measured using glass micropipettes. When the<br />

activity of a single neuron was isolated, a stimulus array that included<br />

100 mM NaCl, 500 mM sucrose, 10 mM HCl, and 20 mM quinine HCl<br />

was applied to the tongue and oral cavity. Taste-responsive units were<br />

found at coordinates relative to obex that correspond to the location of<br />

the rostral NST in mouse neuroanatomical atlases. We recorded the<br />

activity of 18 single cells in B6 mice and 13 cells in 129 mice.<br />

Responses to sucrose were significantly larger in B6 mice. In both<br />

strains, the properties of cells were similar to those reported previously<br />

<strong>for</strong> rats, including breadth-of-tuning values that averaged 0.8. This<br />

represents a large increase over breadth-of-tuning values reported<br />

previously <strong>for</strong> peripheral gustatory neurons in mice, and it suggests that<br />

peripheral fibers with different response profiles converge onto the<br />

same NST neurons in these animals. The larger neural response to<br />

sucrose in B6 mice may underlie their higher preference <strong>for</strong> this<br />

compound in long-term tests. This work was supported by NIH grant<br />

R03 DC005929.<br />

23<br />

91 Poster [ ] Gustatory Processing<br />

THE EFFECT OF FLOW RATE ON THE TEMPORAL<br />

STRUCTURE OF A TASTE RESPONSE RECORDED FROM<br />

THE NUCLEUS OF THE SOLITARY TRACT IN THE RAT<br />

Rugai N. 1, Hallock R.M. 1, Di Lorenzo P.M. 1 1Psychology, State<br />

University of New York at Binghamton, Binghamton, NY<br />

Recent work has suggested that taste quality may be encoded by the<br />

temporal structure of initial 2 sec of response (Di Lorenzo & Victor,<br />

2003) in the nucleus of the solitary tract (NTS). Data from Smith and<br />

Bealer (1975) in the whole chorda tympani nerve (CT) in the hamster<br />

suggest that this interval may also contain in<strong>for</strong>mation about stimulus<br />

flow rate. Here, the effects of flow rate on the temporal structure of<br />

taste responses in the NTS of anesthetized rats were examined. Taste<br />

responses to high (5 ml/s) and low (3.0 ml/s) flow rates of NaCl (0.1 M)<br />

were recorded from taste responsive cells in the rostral nucleus of the<br />

solitary tract (NTS). Each trial consisted of 10 s spontaneous activity,<br />

10 s distilled water, 5 s NaCl, 5 s post-stimulus pause, and a 20 s water<br />

rinse. The inter-stimulus interval was a 2 min. Low and high flow rates<br />

were alternated, and multiple replications were presented. Results<br />

indicated that approximately half of the taste cells also showed<br />

evidence of tactile responsiveness. Approximately half of the cells<br />

showed identical response to NaCl presented at high and low flow rates.<br />

A minority of cells showed an enhanced phasic response related to the<br />

high flow rate, similar to results in the hamster CT. Another subset of<br />

cells showed the opposite effect. Results indicate that flow rate is<br />

encoded within taste responses of NTS cells and that some cells may be<br />

more sensitive to flow rate than others.<br />

Supported by NIDCD grant 1-RO1-DC005219 to PMD<br />

92 Poster [ ] Gustatory Processing<br />

GUSTO-SALIVARY REFLEX CIRCUITS IN RAT BRAINSTEM<br />

Fukami H. 1, Bradley R.M. 1 1Biologic & Materials <strong>Sciences</strong>, University<br />

of Michigan, Ann Arbor, MI<br />

Neural circuits underlying taste-salivary reflexes have been studied<br />

in rat brainstem slices. Parasympathetic neurons innervating the lingual<br />

(von Ebner) and parotid salivary glands were retrogradely labeled by<br />

application of fluorescent Alexofluor dextran to either the<br />

glossopharyngeal or otic ganglion and whole cell patch clamp<br />

recordings made from the labeled neurons. Neurons were first<br />

visualized with fluorescent illumination and once a neuron was selected<br />

<strong>for</strong> recording, it was imaged using differential interference contrast<br />

infrared microscopy. The electrode also contained Lucifer Yellow (LY)<br />

to fill the recorded neuron <strong>for</strong> subsequent identification. After<br />

recording, the LY filled neurons were imaged using a confocal<br />

microscope, reconstructed and morphometrically analyzed. Successful<br />

recordings have been made from 86 neurons. Based on morphometric<br />

measurements neurons innervating the parotid gland are significantly<br />

larger than those innervating the lingual glands. Using various current<br />

injection protocols two major groups of neurons have been defined<br />

based on biophysical and repetitive discharge characteristics. The<br />

majority of the neurons have an action potential discharge pattern that is<br />

modified by the interaction of excitatory and inhibitory synaptic input,<br />

while the minority of neurons transmit the afferent input pattern<br />

unchanged. Since all post-synaptic potentials recorded from the<br />

parasympathetic neurons were a complex mixtures of excitatory and<br />

inhibitory potentials considerable processing of afferent taste input<br />

occurs to determine the output pattern of discharges transmitted by the<br />

secretomotor neurons.<br />

Supported by NIH grant DC000288 to RMB.


93 Poster [ ] Gustatory Processing<br />

ALCOHOL ACTIVATES A SUCROSE-RESPONSIVE<br />

GUSTATORY NEURAL PATHWAY<br />

Lemon C.H. 1, Brasser S.M. 1, Smith D.V. 1 1Dept. of Anatomy &<br />

Neurobiology, University of Tennessee Health Science Center,<br />

Memphis, TN<br />

A strong association exists between the intake of alcohol and sweettasting<br />

substances. The neural mechanisms underlying this relationship<br />

are unknown, although recent data suggest that gustatory factors are<br />

involved. Here, we explored the role of taste receptors and CNS circuits<br />

<strong>for</strong> sugar taste in the gustatory processing of ethanol. Taste responses to<br />

ethanol (3, 5, 10, 15, 25 and 40% v/v) and stimuli of different taste<br />

qualities (e.g., sucrose, NaCl, HCl and quinine-HCl) were recorded<br />

from neurons of the nucleus of the solitary tract in anesthetized rats<br />

prior to and following oral application of the sweet receptor blocker<br />

gurmarin. The magnitude of ethanol-evoked activity was compared<br />

between sucrose-responsive (S , n = 21) and -unresponsive (S , n = 20)<br />

1 0<br />

neurons and the central neural representation of ethanol taste was<br />

explored using multivariate analysis. Ethanol produced robust<br />

concentration-dependent responses in S neurons that were dramatically<br />

1<br />

larger than those in S cells (P's ≤ 0.02). Gurmarin treatment selectively<br />

0<br />

and similarly inhibited ethanol and sucrose responses (P's ≤ 0.01).<br />

Across-neuron patterns of response to ethanol were most similar to<br />

those evoked by sucrose (multiple r = +0.89), becoming increasingly<br />

more so as the ethanol concentration was raised. Results implicate taste<br />

receptors <strong>for</strong> sucrose as candidate receptors <strong>for</strong> ethanol and reveal that<br />

alcohol and sugar taste are represented similarly by activity in the CNS.<br />

These findings have important implications <strong>for</strong> the sensory and hedonic<br />

properties of alcohol. Supported by NIH DC005270 and DC00353.<br />

94 Poster [ ] Gustatory Processing<br />

CHARACTERIZATION OF RAT ORBITOFRONTAL<br />

CORTICAL NEURONS DURING AD LIBITUM DRINKING OF<br />

LIQUID REWARDS.<br />

Gutierrez R. 1, Nicolelis M.A. 1, Simon S.A. 2 1Neurobiology, Duke<br />

University, Durham, NC; 2Anesthesiology, Duke University, Durham,<br />

NC<br />

In gustatory physiology, the orbitofrontal cortex (OFC) contains<br />

neurons that have roles in tongue movements, satiety, and other<br />

motivational outcomes involved in obtaining rewards. There is,<br />

however, a paucity of in<strong>for</strong>mation regarding the nature of the OFC<br />

responses obtained from freely-moving rats licking to obtain a reward.<br />

To this end we have chronically implanted bundles of electrodes in rat<br />

OFC while they were free to drink from a sipper tube water or a sucrose<br />

solution. 172 single unit responses were characterized. A crosscorrelogram<br />

analysis between licking and neural activity revealed that<br />

17% of the responses faithfully followed the licking frequency (~ 7<br />

Hz). Of these oscillating neuronal responses, 24% responded either<br />

more actively or with a different morphology when drinking sucrose<br />

than water, probably reflecting the differential reward values of these<br />

stimuli. To obtain a better understanding of the neuronal activity<br />

involved in the initiation of a licking bout, peri-event histograms were<br />

constructed using the onset of the first lick after at least a 1 sec pause.<br />

Four other morphologically distinct types of responses were identified.<br />

Relative to the initiation of a drinking bout: two types began firing<br />

be<strong>for</strong>e (anticipatory), one type decreased firing and the other increased<br />

firing. In summary, five types of OFC neurons have been identified.<br />

Two, as expected, are related to the anticipation of a reward; others<br />

however are related to licking, and to the nature of the reward.<br />

Supported by NIH DC-01065.<br />

24<br />

95 Poster [ ] Olfactory CNS Physiology and Coding<br />

ONTOGENY OF SENSORY-EVOKED RESPONSES IN RAT<br />

AMYGDALA<br />

Wilson D.A. 1 1Department of Zoology, University of Oklahoma,<br />

Norman, OK<br />

The amygdala plays a critical role in emotion and memory. Recent<br />

work has suggested that adult reactions to emotional and/or stressful<br />

stimuli can be shaped by the effects of early experience on amygdala<br />

functional ontogeny. However, while ontogeny of amygdala neuron<br />

phenotype and neuroananatomy have been described, very few<br />

descriptions of amygdala physiology and function during the postnatal<br />

and adolescent period exist. As a first step toward understanding how<br />

early experiences shape amygdala function, the present study examined<br />

amygdala single-unit activity and responsiveness to sensory input<br />

during postnatal development.<br />

Single-unit and local field potential (LFP) activity were recorded in<br />

amygdala nuclei (primarily basolateral) of urethane-anesthetized, Long-<br />

Evans hooded rats. Rats were aged PN10 to adult (> PN70).<br />

Spontaneous activity, odor-evoked and footshock-evoked activity were<br />

determined <strong>for</strong> each cell using peri-stimulus time histograms.<br />

Spontaneous activity increased and evoked response latency decreased<br />

with age. While both odor-evoked and footshock-evoked responses<br />

could be observed at all ages, the temporal structure of these responses<br />

changed dramatically over the age range tested. Odor-evoked LFP's<br />

showed strong oscillations in the high beta band in adults, however the<br />

primary frequency of these oscillations shifted to lower ranges in<br />

younger animals toward the theta range at PN10. Given the<br />

hypothesized importance of temporal structure in stimulus encoding,<br />

these developmental changes may be indicative of a slow postnatal<br />

emergence of mature sensory processing by the rat amygdala, perhaps<br />

contributing to the sensitivity of this structure to early experiences.<br />

96 Poster [ ] Olfactory CNS Physiology and Coding<br />

FACILITATION OF MAIN OLFACTORY INPUT TO THE<br />

MEDIAL AMYGDALA AND MEDIAL PREOPTIC AREA BY<br />

GONADOTROPIN-RELEASING HORMONE (GNRH).<br />

Blake C. 1, Westberry J. 1, Case G.R. 1, Meredith M. 1 1Neuroscience,<br />

Florida State University, Tallahassee, FL<br />

Sensory signals received during mating activate brain regions along<br />

the vomeronasal pathway and the medial preoptic area. These activated<br />

regions contain cell bodies and fibers of gonadotropin-releasing<br />

hormone (GnRH) neurons, suggesting a possible relationship between<br />

chemosensory input and GnRH. Chemosensory input can be detected<br />

by the vomeronasal organ and/or the main olfactory system. The<br />

consequences of vomeronasal organ removal (VNX) are most apparent<br />

in sexually naïve animals. Experienced, but not inexperienced, male<br />

hamsters can use main olfactory input to maintain mating after VNX,<br />

suggesting that with experience, neural circuits acquire the ability to use<br />

main olfactory in<strong>for</strong>mation as the essential chemosensory input <strong>for</strong><br />

mating. GnRH has also been shown to restore mating behavior in naïve<br />

VNX male hamsters. GnRH may facilitate transfer of olfactory<br />

in<strong>for</strong>mation to the medial amygdala and medial preoptic area (MPOA)<br />

in naïve intact and VNX males. Electrical stimulation of the main<br />

olfactory bulb activates neurons in anterior and posterior medial<br />

amygdala and increases expression of FRAs (Fos-related antigens)<br />

protein. Initial experiments indicate an icv injection of GnRH into the<br />

lateral ventricle increases activation in medial amygdala and medial<br />

preoptic area. These experiments examine the effect of GnRH on<br />

chemosensory in<strong>for</strong>mation transfer via the main olfactory pathway to<br />

these regions that are involved in mating behavior driven by either<br />

chemosensory pathway. Supported by DC-005813 from NIDCD.


97 Poster [ ] Olfactory CNS Physiology and Coding<br />

ODOR RESPONSES OF CEREBRAL LOBE NEURONS<br />

Nikonov A.A. 1, Caprio J. 1 1Biological <strong>Sciences</strong>, Louisiana State<br />

University, Baton Rouge, LA<br />

Functional in<strong>for</strong>mation as to how odors are processed at more central<br />

neural levels than the olfactory bulb (OB) is fragmentary at best in<br />

mammals and nonexistent in teleosts. The long-term goal is to<br />

determine similarities and differences in how odor in<strong>for</strong>mation arriving<br />

via the olfactory tracts is processed within the cerebral lobes (CL) of<br />

the channel catfish compared to that in the OB. Odor-responsive<br />

neurons were located in caudo-medial and lateral regions of the CL as<br />

predicted from previous <strong>for</strong>ebrain anatomical investigations (Finger,<br />

1975; Bass, 1981). Recordings were per<strong>for</strong>med in vivo with metal-filled<br />

glass microelectrodes. Tested were bile salts (taurocholic and<br />

taurolithocholic acids) and amino acids (L-methionine and L-arginine),<br />

effective odor stimuli in channel catfish (Nikonov & Caprio, 2001).<br />

Sixty-four CL neurons responded to the odor stimulation; 45 were<br />

excited and 19 were suppressed (12 by amino acids and 7 by bile salts).<br />

Fifteen units located medially in the CL were excited by bile salts and<br />

were unresponsive to amino acids. Twenty-six of the 30 excitatory units<br />

located more laterally responded to 0.1-100µM amino acid; 4 additional<br />

units within this region along the OB midline responded to bile salts.<br />

The medial-lateral distinction between excitatory responses to bile salts<br />

and amino acids reflects a similar topographical organization within the<br />

OB. Ongoing experiments will determine if an odotopic map as<br />

observed in the OB (Nikonov and Caprio, 2001) exists or is altered<br />

within the CL and whether the response properties of CL neurons are<br />

different from those of OB neurons to the same odors.<br />

Supported by NSF IBN-0314970 and NIH DC-03792<br />

98 Poster [ ] Olfactory CNS Physiology and Coding<br />

ODOR REPONSIVENESS OF THE OLFACTORY CORTEX<br />

AND RESIDUAL TYROSINE HYDROXYLASE ACTIVITY IN<br />

THE OLFACTORY BULB IN THE CNGA2 KNOCKOUT<br />

MOUSE<br />

Lin W. 1, Restrepo D. 1 1Cell and Developmental Biology, Neuroscience<br />

Program and Rocky Mountain Taste and Smell Center, University of<br />

Colorado Health <strong>Sciences</strong> Center, Denver, CO<br />

Mice deficient <strong>for</strong> subunit A2 of the cyclic nucleotide-gated (CNG)<br />

channel respond to select odorants (see Restrepo et al abstract in this<br />

meeting of AChemS). We examined the presence of alternate<br />

transduction pathways and the potential role of necklace glomeruli,<br />

which receive axons from neurons expressing the cGMP pathway<br />

(Juilfs et al 1997, PNAS 94:3388-95) and remain active in CNGA2 KO<br />

mice (Baker et al 1999, J Neurosci. 19:9313-21). Immunoreactivity to<br />

the cGMP-stimulated phosphodiesterase (PDE2), tyrosine hydroxylase<br />

(TH) and odor-induced Fos was used to visualize necklace glomeruli,<br />

levels of glomerular activity and activation respectively. We found that<br />

the TH expression in juxtaglomerular cells surrounding PDE2-positive<br />

necklace glomeruli is dependent on afferent input as unilateral naris<br />

occlusion dramatically reduced the TH expression in CNGA2 KO mice.<br />

Importantly, exposure CNGA2 KO mice to a mixture of putative<br />

pheromones (2-heptanone and 2, 5-dimethylpyrazine) increased odorevoked<br />

Fos induction in juxtaglomerular cells of some necklace and<br />

regular glomeruli. Further, we observed Fos expression in a<br />

substantially larger number of cells in both the anterior olfactory<br />

nucleus (AON) and piri<strong>for</strong>m cortex in CNGA2 knockout mice<br />

compared to the same regions from mice exposed to fresh air. These<br />

experiments provide evidence <strong>for</strong> alternate transduction pathways in the<br />

main olfactory system in mice.<br />

This work was supported by NIH grants DC00566, DC04657,<br />

DC006070 (DR) and DC0043(WL).<br />

25<br />

99 Poster [ ] Olfactory CNS Physiology and Coding<br />

SYNAPTIC MECHANISMS CONTRIBUTING TO OLFACTORY<br />

CORTICAL ADAPTATION<br />

Best A.R. 1, Wilson D.A. 1 1Zoology, University of Oklahoma, Norman,<br />

OK<br />

Anterior piri<strong>for</strong>m cortex (aPCX) neurons rapidly filter repetitive odor<br />

stimuli despite relatively maintained input from mitral cells. This<br />

cortical adaptation is correlated with short-term depression of afferent<br />

synapses, in vivo. The purpose of this study was to elucidate<br />

mechanisms underlying this non-associative plasticity using in vivo and<br />

in vitro preparations and to determine its role in cortical odor<br />

adaptation. We have previously described our in vitro stimulation<br />

results. In particular, we have discovered that the longer lasting portion<br />

(~120 s) of the in vitro short-term synaptic depression can be blocked<br />

pharmacologically. Following 50 s of simulated odor stimulation, either<br />

the metabotropic glutamate II/III antagonist CPPG (500 µM) or the βadrenergic<br />

receptor agonist isoproterenol (20 µM) can decrease the<br />

duration of synaptic depression. More recently we have discovered that,<br />

in vivo, both adaptation of odor responses in the β (15-35 Hz) spectral<br />

range and the associated synaptic depression following a 50 s exposure<br />

to a mixture of odorants can also be blocked by intracortical infusion of<br />

CPPG (2.5 mM). Currently, we are extending this work to behavioral<br />

habituation using heart rate orienting response habituation as a measure<br />

of cortical adaptation. Prior to adaptation of the heart rate orienting<br />

response, 3 µl of 2.5 mM CPPG or vehicle will be infused into aPCX<br />

bilaterally. Results from the behavioral paradigm will assist in<br />

clarifying the independent roles of receptor adaptation and cortical<br />

adaptation in behavioral habituation to odors. Funded by NICHD.<br />

100 Slide [ ] Olfactory CNS Physiology and Coding<br />

RAPID ODOR CHANGE AND BRAIN ELECTRICAL<br />

ACTIVITY<br />

Lorig T.S. 1, Rigdon M. 1, Poor A. 1 1Washington and Lee University,<br />

Lexington, VA<br />

Some odors produce increased activity in areas of the brain often<br />

associated with language even though the task demands do not<br />

explicitly involve language. Visual and auditory stimuli which change<br />

rapidly tend to produce more left temporal lobe activity than stimuli<br />

that change slowly. If odor mixtures are transduced in such a way that<br />

each component produces high temporal overlap with the other<br />

components, this should effectively be construed as rapid stimulus<br />

change and lead to more left temporal lobe responses. To test this<br />

hypothesis, 10 undergraduate students served as subjects in a<br />

chemosensory event-related potential (CSERP) study. Three odors,<br />

phenethyl alcohol, vanillin, and citral were presented in binary pairs via<br />

a constant flow olfactometer and stimulus administration was triggered<br />

by subjects' inhalation. The stimuli were presented in two different<br />

ways. Given the members of the binary odor pair, A and B, one<br />

condition (Slow) required subjects to smell a sequence of A followed<br />

by B. Both A and B lasted 300 msec and were counterbalanced. In the<br />

Fast condition, the sequence was A,B,A,B where each stimulus lasted<br />

150 msec and was counterbalanced. In both cases, each odor lasted a<br />

total of 300msec and the total stimulus administration lasted 600msec.<br />

CSERP data were collected from 80 scalp locations. Analysis revealed<br />

that brain activity differed between the two types of stimulus<br />

administration. Subsequent analysis with LORETA indicated that the<br />

left temporal lobe was more active in the Fast condition supporting the<br />

hypothesis that temporal overlap in odor mixtures leads to more left<br />

temporal lobe activity.


101 Poster [ ] Olfactory CNS Physiology and Coding<br />

PET ACTIVATIONS DURING SMELLING OF<br />

ANDROSTENONE: OSMICS VERSUS ANOSMICS<br />

Boyle J.A. 1, Zatorre R. 1, Pause B. 2, Jones Gotman M. 1 1McGill<br />

University, Montreal, Quebec, Canada; 2Christian-Albrechts-<br />

University, Kiel, Germany<br />

We used positron emission tomography to evaluate the neural<br />

correlates involved in perception of the endogenous steroid<br />

androstenone. Our aim was to investigate whether differences exist in<br />

brain activations between subjects with high sensitivity (low thresholds)<br />

<strong>for</strong> androstenone and those with a specific anosmia. We screened 87<br />

subjects using a staircase procedure to establish individual thresholds.<br />

Twelve subjects unable to detect the steroid in its crystal <strong>for</strong>m were<br />

selected <strong>for</strong> an anosmic group, and 12 subjects capable of consistently<br />

detecting a 1.4 x 10-7 M solution of androstenone comprised an osmic<br />

group. Subjects were scanned during passive smelling of phenyl ethyl<br />

alcohol (PEA), pyridine, androstenone (AND) or the diluant (<strong>for</strong><br />

baseline). Our first analysis compared brain responses of the two groups<br />

to two of these stimuli: PEA and AND versus baseline. Activations <strong>for</strong><br />

PEA were observed in olfactory regions, including orbitofrontal and<br />

piri<strong>for</strong>m cortex, in both subject groups. However, osmics and anosmics<br />

differed in their response to AND. Osmics showed more activity in the<br />

dorsomedial area of the left parietal cortex and had no significant<br />

olfactory activations, whereas anosmics showed more frontal<br />

activations in the right dorsomedial and orbitofrontal regions. The<br />

results <strong>for</strong> PEA are consistent with other studies of olfactory<br />

processing, showing activations in olfactory cortices, whereas the AND<br />

results suggest differential function in the human brain related to<br />

individual sensitivity to androstenone. Implications of the parietal lobe<br />

activations will be discussed. Supported by Canadian Institutes of<br />

Health Research<br />

102 Poster [ ] Olfactory CNS Physiology and Coding<br />

FUNCTIONAL NEUROIMAGING OF ODOR IMAGERY<br />

Djordjevic J. 1, Zatorre R. 1, Petrides M. 1, Boyle J. 1, Jones Gotman M. 1<br />

1Montreal Neurological Institute, Montreal, Quebec, Canada<br />

We used positron emission tomography to investigate brain regions<br />

associated with odor imagery. We compared changes in brain activation<br />

during active imagination of odors versus those during nonspecific<br />

expectation of olfactory stimuli, and during experience of physically<br />

presented versus imagined odors. Sixty-seven healthy volunteers were<br />

screened <strong>for</strong> their odor imagery (with a paradigm developed in a<br />

previous study), and 12 of them, selected as “good odor imagers”,<br />

participated in the neuroimaging study. Comparison of odor imagery<br />

with general expectation of odors revealed activation in the left primary<br />

olfactory cortex (POC) including piri<strong>for</strong>m cortex, as well as in the<br />

insula bilaterally, and the left posterior orbitofrontal cortex (OFC). On<br />

the other hand, comparison of odor perception with an odorless baseline<br />

resulted in increased activation bilaterally in the piri<strong>for</strong>m and insular<br />

cortex and in the right posterior OFC. Results of a conjunction analysis<br />

revealing regions of common activation during odor perception and<br />

imagery were convergent with the results obtained by the subtraction<br />

method: left piri<strong>for</strong>m cortex and bilateral insula were activated both<br />

when smells were perceived and imagined. Findings also suggested a<br />

differential role of the right versus left posterior OFC in perceptual<br />

versus imaginal olfactory processing: whereas the right OFC was<br />

activated during odor perception, the same region of the left hemisphere<br />

was activated during odor imagery. Overall, the findings indicated that<br />

neural networks engaged during odor perception and imagery partially<br />

overlap.<br />

Supported by Canadian Institutes of Health Research<br />

26<br />

103 Slide [ ] Olfactory CNS Physiology and Coding<br />

NEURAL SUBSTRATES UNDERLYING MENTAL IMAGERY<br />

OF PLEASANT AND UNPLEASANT SMELLS<br />

Bensafi M. 1, Porter J.A. 2, Khan R.M. 1, Mainland J. 1, Johnson B. 3,<br />

Zelano C. 4, Sobel N. 1 1Neuroscience, University of Cali<strong>for</strong>nia,<br />

Berkeley, Berkeley, CA; 2Psychology, University of Cali<strong>for</strong>nia,<br />

Berkeley, Berkeley, CA; 3Bioengineering, University of Cali<strong>for</strong>nia,<br />

Berkeley, Berkeley, CA; 4Biophysics, UC Berkeley, Berkeley, CA<br />

Patterns of neural activity in orbito-frontal cortex (OFC) reflect odor<br />

valence: the medial OFC responds preferentially to pleasant odors,<br />

whereas the lateral OFC responds preferentially to unpleasant odors.<br />

Whether this dissociation exists during odor imagery in the absence of<br />

any odorants is unknown. Here, we approached this question using<br />

fMRI (4T). Sixteen subjects (8f) were tested in an event-related design<br />

with 5 trial types, each repeated 22 times (TR=500ms, ISI=30sec). In 3<br />

sensory trial types, a pleasant odor (strawberry), an unpleasant odor<br />

(rotten eggs) and a non-odorized control were presented by air dilution<br />

olfactometer. In 2 imagery trial types, no odorants were presented, but<br />

subjects were asked to imagine the same smells (strawberry, rotten<br />

eggs). Replicating previous reports, initial analysis revealed a<br />

dissociation during perception whereby strawberry induced a significant<br />

activation in medial OFC, whereas rotten eggs induced an increase in<br />

lateral OFC (p


105 Poster [ ] Unicellular <strong>Chemoreception</strong><br />

GPI ANCHORED RECEPTORS IN CHEMOSENSORY<br />

TRANSDUCTION<br />

Weeraratne S.D. 1, Valentine M. 1, Yano J. 1, Van Houten J. 1 1Biology,<br />

University of Vermont, Burlington, VT<br />

Glycosyl phosphatidylinositol (GPI) anchored proteins are surface<br />

proteins that are tethered to the cell surface. They serve as receptors of<br />

ligands in axon guidance and T cell activation among other functions.<br />

In Paramecium they function in chemoreception. We have shown<br />

through disruption of anchoring using an antisense expression to PIG-A<br />

(an enzyme in the fist step in anchor synthesis) that chemoresponses to<br />

folate and glutamate are almost eliminated. Disruption of the last<br />

enzyme of anchor synthesis with an RNAi expression vector also<br />

disrupted chemoresponse, but not as profoundly. The genes <strong>for</strong> the first<br />

and last enzymes in the pathway were cloned using homology PCR, and<br />

with the sequencing of the Paramecium genome, we have identified the<br />

sequences of the other pathway enzymes. We have examined the<br />

distribution of the GPI anchored proteins on the cell surface and found<br />

that they are not evenly distributed across the surface, but rather they<br />

cluster near the bases of the cilia, where another chemosensory<br />

transduction component (the plasma membrane calcium pump- PMCA)<br />

is also found. We show this distribution through confocal and<br />

deconvolution microscopy using antibodies to the folate chemoreceptor,<br />

5´ nucleotidase, GPI anchored surface antigens, tubulin and the<br />

PMCAs. The Paramecium genome project has provided us with<br />

sequences <strong>for</strong> putative GPI anchored folate receptors and our<br />

preliminary results using RNAi-feeding support the hypothesis that at<br />

least one of these proteins functions in folate chemoresponse.<br />

Supported by NIH DC00721, GM59988, the Vemont Cancer Center.<br />

106 Poster [ ] Unicellular <strong>Chemoreception</strong><br />

LIPID RAFTS IN CHEMOSENSORY TRANSDUCTION<br />

Chandran S. 1, Ray K. 1, Yano J. 1, Van Houten J. 1 1Biology, University of<br />

Vermont, Burlington, VT<br />

Lipid Rafts are lipid domains high in glycosphingolipids and<br />

cholesterol that serve as plat<strong>for</strong>ms <strong>for</strong> organizing signal transduction<br />

components such as GPI anchored proteins, G proteins (large and<br />

small), serpentine receptors, receptor and non-receptor tyrosine kinases.<br />

Proteins in lipid rafts are insoluble in cold Triton X-100 and are found<br />

in low density fractions when the Triton insoluble fractions are<br />

separated in sucrose or Optiprep gradients. We have found that<br />

Paramecium surface membranes can be separated in Optiprep gradients<br />

into fractions by density, and correlated with ganglioside and<br />

cholesterol levels. Proteins such as the putative folate chemoreceptor<br />

(and other GPI anchored proteins such as 5´ nucleotidase and surface<br />

antigens), G-beta, and the plasma membrane calcium pumps separate<br />

into the light density fractions, supporting the existence of lipid rafts<br />

that organize these proteins in signaling. When we remove cholesterol<br />

from the membranes of living cells using methyl-beta-cyclo-dextrin, we<br />

find that chemoresponse to folate is significantly reduced, implying<br />

that raft organization is important <strong>for</strong> chemoresponse. We are also<br />

using cholera toxin to localize gangliosides as markers of rafts on the<br />

cell surface, and asking whether removal of cholesterol changes the<br />

distribution of signaling proteins in Optiprep gradients and on the cell<br />

surface. Supported by GM59988, DC 00721, and Vermont Cancer<br />

Center.<br />

27<br />

107 Poster [ ] Unicellular <strong>Chemoreception</strong><br />

PLASMA MEMBRANE CALCIUM PUMPS FUNCTIONING IN<br />

CHEMICAL SENSING<br />

Yano J. 1, Zhukovskaya M. 2, Preston R. 3, Pan Y. 1, Keiser M. 1, Van<br />

Houten J. 1 1Biology, University of Vermont, Burlington, VT;<br />

2Pharmacology and Physiology, Drexel University, PA, PA;<br />

3Pharmacology and Physiology, Drexel University, Philadelphia, PA<br />

The Plasma Membrane Calcium Pumps (PMCAs) play a role in<br />

sensory transduction in Paramecium by sustaining the hyperpolarizing<br />

conductance that is initiated by a K conductance upon stimulus<br />

application. We identified 4 PMCA iso<strong>for</strong>ms through homology PCR,<br />

and found that 3 were expressed. The Paramecium genome project<br />

shows that there are up to 4 more related sequences that probably code<br />

<strong>for</strong> additional PMCAs. To explore the roles of these PMCAs we have<br />

begun to over-express them with and without epitope tags to localize<br />

them on the cell surface, and found that over expression disrupts<br />

calcium homeostasis in the cell. The cells´ swimming patterns are<br />

dependent upon the intraciliary calcium levels, and over-expression of<br />

the iso<strong>for</strong>ms 2 and 3 cause cells to swim backward <strong>for</strong> long periods of<br />

time, consistent with high intraciliary calcium. We have also found that<br />

the calcium conductance of the voltage gated calcium channels is<br />

significantly changed in cells over-expressing the iso<strong>for</strong>ms 2 and 3,<br />

consistent with abnormally high intracellular calcium and defective<br />

calcium homeostasis. This has allowed us to systematically over<br />

express the iso<strong>for</strong>ms and ask which chemoresponse is affected. We<br />

find that iso<strong>for</strong>m 2 over-expression primarily affects folate<br />

chemoresponse, while iso<strong>for</strong>m 3 over expression affects chemoreponse<br />

to folate, glutamate, acetate, all of which function through different<br />

signal transduction pathways, but all of which we believe to include the<br />

PMCAs. Supported by DC 00721, Vermont Cancer Center.<br />

108 Poster [ ] Unicellular <strong>Chemoreception</strong><br />

THE RYANODINE RECEPTOR ANTAGONIST DANTROLENE<br />

ALTERS SWIMMING BEHAVIOR AND CAUSES MORTALITY<br />

IN PARAMECIUM TETRAURELIA<br />

Green M.H. 1, Brown H.J. 1, Bell W.E. 1 1Biology, Virginia Military<br />

Institute, Lexington, VA<br />

Paramecium tetraurelia are unicellular eukaryotes with excitable<br />

ciliated membranes that use calcium as a regulator of membrane<br />

potential. Several critical Paramecium functions are known to be<br />

mediated by calcium such as chemoresponse and exocytosis. Because<br />

caffeine and 4-chloro-m-cresol stimulation of Paramecium results in an<br />

increase in intracellular calcium concentration ( Klauke, et al., 2000, J.<br />

Mem. Biol. 174:141-156), it is possible that a calcium channel similar<br />

to the ryanodine receptors found in skeletal muscle may be involved in<br />

calcium mobilization. We examine the effects of the ryanodine<br />

receptor antagonist, dantrolene, on Paramecium tetraurelia. Dantrolene<br />

was toxic when applied in doses normally used in mammalian systems<br />

in the presence of magnesium. It is possible that magnesium, which also<br />

is a ryanodine receptor inhibitor, amplifies the effect of dantrolene on<br />

ryanodine-like receptors in Paramecium resulting in toxicity. Also of<br />

interest is the lack of protection from dantrolene-mediated toxicity in<br />

the mutant eccentric, which has a defective magnesium conductance.<br />

Initial experiments indicate that dantrolene does not affect the ability of<br />

caffeine to increase intracellular calcium levels. Dantrolene does<br />

however, significantly reduce the backward swimming time of<br />

Paramecium exposed to depolarizing concentrations of potassium.<br />

Chemosensory behavior in T-Maze assays to the attractants biotin and<br />

acetate was not altered by dantrolene treatment in potassium solutions.<br />

This work was supported by grant J-651 from the Jeffress Memorial<br />

Trust and a VMI research Grant in Aid.


109 Poster [ ] Clinical Evaluation and Consumer Research<br />

COMPUTERIZED HISTORY OF OLFACTORY<br />

DYSFUNCTION<br />

Cornelia H. 1, Landis B.N. 2, Frasnelli J. 1, Hummel T. 1 1ORL, University<br />

of Dresden Medical School, Dresden, Germany; 2ORL, University<br />

Hospital of Geneva, Geneva, Switzerland<br />

The patient´s/subject´s history is an integral part of investigations on<br />

human olfactory sensitivity. Numerous ways are being employed to<br />

obtain in<strong>for</strong>mation from subjects/patients. Based on Filemaker<br />

Developer 6.0 (R), the current exercise aimed to produce a relatively<br />

short, computerized questionnaire which allows subjects/patients to<br />

enter their data directly into a database in a controlled fashion. Further,<br />

a separate section is availbale <strong>for</strong> the examiner to enter data with regard<br />

to drug intake, qualitative olfactory dysfunction, results of<br />

chemosensory testing, results from nasal endoscopy, treatment etc. A<br />

different set of questions is available <strong>for</strong> return visits. In addition, the<br />

present approach combines this database with software <strong>for</strong> the<br />

computerized testing of olfactory thresholds (triple-<strong>for</strong>ced choice,<br />

single staircase), odor discrimination (triple <strong>for</strong>ced choice, 16 triplets)<br />

and odor identification (mutiple <strong>for</strong>ced choice, 16 items). In a clinical<br />

context, the database allows to create a quick overview across changes<br />

of olfactory function over time. It is hoped that this approach will<br />

eventually lead to more complete in<strong>for</strong>mation from subjects/patients.<br />

Further, it is hoped that this approach will help to standardise<br />

in<strong>for</strong>mation that is obtained at different centers.<br />

110 Poster [ ] Clinical Evaluation and Consumer Research<br />

CLINICAL TEST OF OLFACTION BASED UPON A MEMS-<br />

MICROVALVE OLFACTOMETER.<br />

Hastings L. 1, Wilson T. 1 1Osmic Enterprises, Inc., Cincinnati, OH<br />

We describe the development of a clinical test <strong>for</strong> the assessment of<br />

olfactory function which incorporates state-of-the-art Micro-Electro-<br />

Mechanical Systems (MEMS) microvalves. These minature, discrete<br />

devices <strong>for</strong>m the basis of the MEMS olfactometer, a relatively small,<br />

inexpensive device capable of producing 40 or more discrete odor<br />

stimuli. Moreover, the MEMS olfactometer will have the capacity <strong>for</strong><br />

generating the stimuli necessary <strong>for</strong> a large number of trials (>100).<br />

This capability can be replenished by replacing inexpensive odor<br />

cartridges. The initial test selected <strong>for</strong> development is an odor<br />

identification test, the paradigm most frequently used to assess<br />

olfactory function. The test, called the OLFACT (OLfactory Function<br />

Assessment by Computerized Testing) will be enhanced by inclusion of<br />

pictures, along with words, in the description of test item choices. The<br />

presentation of the test items, along with scoring of the test, and<br />

recording of all pertinent data will be totally computerized. Norms <strong>for</strong><br />

the test will be developed and compared with standardized norms<br />

already available <strong>for</strong> other commercial tests. Finally, the test will be<br />

enhanced to run on the WWW as a Web application. Once a MEMS<br />

olfactometer has been attached to a computer linked to the WWW,<br />

administration of the tests can be authorized over the Web. Anonymous<br />

data can also be collected via the Web <strong>for</strong> inclusion into a centralized<br />

database. This test will provide a comprehensive and sensitive system<br />

<strong>for</strong> evaluating the sense of smell at a lower cost while providing greater<br />

utility than current tests. Supported by NIH grant DC 06369<br />

28<br />

111 Poster [ ] Clinical Evaluation and Consumer Research<br />

HEART RATE CHANGES DURING ODORANT<br />

ADMINISTRATION: PROMOTION OF "COOL-DOWN" AND<br />

RECOVERY IN COLLEGE ATHLETES<br />

Smith J. 1, Raudenbush B. 1 1Psychology, Wheeling Jesuit University,<br />

Wheeling, WV<br />

An often under-addressed aspect of athletic training, and even casual<br />

exercise, is the proper amount of time <strong>for</strong> "cool-down" and recovery.<br />

However, when an ample recovery period is not available, the<br />

likelihood of injury and overtraining increases while athletic<br />

per<strong>for</strong>mance decreases. Previous research has shown that odorants can<br />

affect one's mood, motivation, and task per<strong>for</strong>mance. Moreover,<br />

peppermint odor is linked to enhanced athletic per<strong>for</strong>mance, while<br />

jasmine odor is a proven sleep aid. These unique odorant characteristics<br />

led to their inclusion within the present experiment in an attempt to<br />

determine whether jasmine and peppermint odors can enhance athletic<br />

recovery. In a within-subjects design, twenty athletes per<strong>for</strong>med a<br />

modified version of the Bruce Stress Test Protocol on a treadmill <strong>for</strong> 15<br />

minutes and then completed push-ups until exhaustion. Following 10<br />

minutes of "cool-down" stretching in a peppermint, jasmine, or no-odor<br />

condition, physiological data were recorded and the participant<br />

completed questionnaires related to workload demands and mood. In<br />

addition, level of vigor was rated over the following twelve hours.<br />

Jasmine odor significantly reduced athletes' heart rate following the<br />

"cool-down" period compared to the non-odorized control condition.<br />

Such a finding supports the hypothesis that odorants may have a<br />

substantial role in naturally and safely expediting recovery from<br />

physical exertion. This study was funded by a grant from NASA to B.<br />

Raudenbush.<br />

112 Poster [ ] Clinical Evaluation and Consumer Research<br />

EFFECTS OF PEPPERMINT ODOR ADMINISTRATION ON<br />

AUGMENTING BASKETBALL PERFORMANCE DURING<br />

GAME PLAY<br />

Raudenbush B. 1, Smith J. 1, Graham K. 1, Mc Cune A. 2 1Psychology,<br />

Wheeling Jesuit University, Wheeling, WV; 2Physical Therapy,<br />

Wheeling Jesuit University, Wheeling, WV<br />

Previous research indicates that inhaling peppermint odor prior to<br />

and during athletic activity increases strength, speed, and endurance. It<br />

has also been found to reduce fatigue, perceived ef<strong>for</strong>t, and perceived<br />

frustration, and increase levels of vigor and motivation. However,<br />

assessment of peppermint odor efficacy has yet to be per<strong>for</strong>med during<br />

actual physical game play. The present study was designed to assess<br />

whether the degree to which athletes inhale peppermint odor affects<br />

such aspects as motivation, energy, fatigue, reaction time, confidence,<br />

and per<strong>for</strong>mance during the course of a basketball season. Male and<br />

female Division II basketball players were provided with a peppermint<br />

inhaler (Peak Per<strong>for</strong>manceTM Sports InhalerTM) <strong>for</strong> use during<br />

practice and game play. Level of inhalant use constituted group<br />

composition <strong>for</strong> data analysis. Higher levels of inhalant use were<br />

associated with increased motivation, energy, speed, alertness, reaction<br />

time, confidence, and strength. Levels of fatigue and frustration were<br />

lower in the high-use group. In addition, athletes' ratings of their<br />

competitive advantage over opponents and ratings of overall<br />

per<strong>for</strong>mance were enhanced. Implications are particularly salient in<br />

regards to augmenting a variety of factors related to athletic<br />

per<strong>for</strong>mance using an all-natural, non-pharmacological ergogenic aide.


113 Poster [ ] Clinical Evaluation and Consumer Research<br />

EFFECTS OF BEVERAGE FLAVOR ON ATHLETIC<br />

PERFORMANCE, MOOD, AND WORKLOAD<br />

Schuler A. 1, Rawson A. 1, Raudenbush B. 1 1Psychology, Wheeling Jesuit<br />

University, Wheeling, WV<br />

Previous research indicates that the administration of peppermint<br />

odor can augment athletic per<strong>for</strong>mance and mood, and decrease<br />

workload demands. The present study extended those findings by<br />

evaluating athletic per<strong>for</strong>mance and physiological changes during the<br />

administration of flavored beverages. Utilizing a within-subjects design,<br />

athletes per<strong>for</strong>med a 15-min modified treadmill stress test. At 3-min<br />

intervals, 50 mL of beverage (peppermint water, unadulterated water, or<br />

Gaterade sports drink) were consumed. In the control condition, no<br />

beverage was consumed. Pre- and post-testing physiological<br />

measurements were taken (blood pressure, pulse, oxygen<br />

concentration). In addition, ratings of mood (via the Profile of Mood<br />

States) and workload (via the NASA Task Load Index) were completed.<br />

No physiological changes were noted, however, both the peppermint<br />

and Gatorade sports drink conditions lead to greater ratings of personal<br />

per<strong>for</strong>mance and increased mood. These results provide additional<br />

support <strong>for</strong> the implementation of non-pharmacological methods to<br />

increase an athlete's per<strong>for</strong>mance and mood during exercise and/or<br />

competition. This study was funded by a grant from NASA to B.<br />

Raudenbush.<br />

114 Poster [ ] Clinical Evaluation and Consumer Research<br />

OLFACTORY LOSS FOLLOWING INTRANASAL ZINC<br />

GLUCONATE<br />

Jafek B.W. 1, Linschoten M.R. 2 1Otolaryngology, University of<br />

Colorado Health <strong>Sciences</strong> Center, Denver, CO; 2Otolaryngology-Head<br />

& Neck Surgery, University of Colorado Health <strong>Sciences</strong> Center,<br />

Denver, CO<br />

Beneficial zinc absorption takes place via enteral, parenteral, or<br />

cutaneous routes. Direct application to the olfactory epithelium,<br />

however, has been reported to cause loss of smell. Intranasal zinc<br />

gluconate has recently been recommended as a treatment <strong>for</strong> the<br />

common cold. Severe posttreatment hyposmia and anosmia have been<br />

observed. The case report of a typical patient will be presented and<br />

analyzed in detail, followed by a series of patients with severe<br />

hyposmia or anosmia following the use of intranasal zinc gluconate in<br />

both of the two currently available commercial products. While<br />

interindividual variation in drug response and drug effect is apparent,<br />

the loss may be permanent. The mechanism of olfactory loss is thought<br />

to be the direct effect of the divalent zinc ion on the olfactory receptor<br />

cell.<br />

29<br />

115 Poster [ ] Clinical Evaluation and Consumer Research<br />

CHEMOSENSORY CHANGES IN ESTROGEN RECEPTOR<br />

POSITIVE BREAST CARCINOMA: A CASE REPORT<br />

Bailey M.L. 1, Hirsch A.R. 2 1Lake Erie College of Osteopathic<br />

Medicine, Erie, PA; 2The Smell & Taste Treatment and Research<br />

Foundation, Chicago, IL<br />

ABSTRACT: Despite reduction in appetite in patients with cancer,<br />

physicians rarely assess chemosensory function. A patient with<br />

estrogen receptor positive (ER+) breast carcinoma presents with<br />

chemosensory complaints.<br />

CASE PRESENTATION: A fifty-five year old female, presented<br />

with seven months of taste distortion. She initially noted a<br />

phantageusia of “Windex” which changed to a soapy, metallic<br />

sensation, localized to the posterior tongue and upper lip. Drinking<br />

water improved the phantageusia, and consuming sweets, milk, vinegar<br />

or acidic foods worsened the phantageusia. Taste threshold testing<br />

revealed hypogeusia to bitter, salt, and sour. Six weeks after<br />

evaluation, she underwent modified radical mastectomy <strong>for</strong> ER (+)<br />

breast carcinoma. Immediately thereafter, the phantageusia resolved<br />

and her perception of sweet, salty, and sour taste intensified. Repeat<br />

chemosensory testing one month later on no medications, prior to any<br />

chemotherapy, demonstrated marked improvement <strong>for</strong> all taste<br />

modalities except salt, which remained unchanged.<br />

DISCUSSION: Possible origins <strong>for</strong> gustatory deficits include remote<br />

effects of carcinoma, counter-stimulation from circulating tumorinduced<br />

hematogenous tastes, tumor released hematogenous synergistic<br />

tastes, taste bud regeneration inhibition, zinc deficiency, elevated<br />

calcium or lactate levels, dry mucous membranes, and lack of interest<br />

and motivation. Chemosensory evaluation is warranted in patients who<br />

present with ER (+) breast carcinoma.<br />

116 Poster [ ] Clinical Evaluation and Consumer Research<br />

ODORANTS AT THE WORLD TRADE CENTER DISASTER<br />

SITE: ANALYSIS AND PSYCHOLOGICAL IMPACT<br />

Preti G. 1, Opiekun R. 1, Smeets M. 1, Fatsis S. 2, Dalton P. 1 1Monell<br />

Chemical Senses Center, Philadelphia, PA; 2The Wall Street Journal,<br />

Washington, DC<br />

Odors present during a traumatic event may become associated with<br />

internal and external aspects of the experience and in turn, can trigger<br />

an emotional or stress response when encountered at a later time.<br />

Knowledge of the odor-causing molecules present at disaster sites is<br />

central to developing a synthetic odor-mimic, which can be used to<br />

prospectively educate rescue workers and to desensitize those who have<br />

already developed odor-stress associations. The distinct and pervasive<br />

odors lingering in the vicinity of lower Manhattan <strong>for</strong> weeks following<br />

9-11 were prime candidates <strong>for</strong> eliciting odor-mediated “flash-backs”<br />

among worker and residents. To identify and describe the quality of the<br />

odorants contributing to this unique, but complex, smell, four of the<br />

authors rated the sensory attributes and collected air samples using<br />

Tedlar bags and SPME “field units” at two sites surrounding the<br />

collapsed World Trade Center towers. Using GC-Olfactometry, the<br />

individuals who experienced the odors at the site evaluated the odorants<br />

as they emerged from the GC, being particularly careful to identify<br />

odorants that could be linked to the characteristic odor of the site. Many<br />

of the characteristic odorants were linked to specific compounds (eg<br />

“smokey” from guiacol; “sour/musty” from C7 to C9 acids; ) or small<br />

groups of compounds (“musty/burnt” and irritating/vinegar-like from a<br />

combination of butyrolactone/benzycyanide/triethylenediamine) which<br />

are commercially available and may be used to reconstitute the<br />

characteristic odor.<br />

Supported by NIH R01 DC 03704 to PD


117 Poster [ ] Clinical Evaluation and Consumer Research<br />

ODOR CONDITIONING AND THE STRESS RESPONSE<br />

Maute C. 1, Dalton P. 1, Michele G. 1 1Monell Chemical Senses Center,<br />

Philadelphia, PA<br />

The present study evaluated the degree to which an emotional state<br />

(i.e. stress or relaxation) initially experienced in the presence of a novel<br />

odor could later be elicited by the odor alone. Evaluations of both<br />

subjective (self-reported stress, health symptoms and mood) and<br />

objective (salivary cortisol) indices of stress demonstrated the efficacy<br />

of the Trier Social Stress Test (TSST) as a stress manipulation. Further,<br />

salivary cortisol levels revealed that subjects who were administered the<br />

(TSST) in the presence of a novel odor were more stressed upon reexposure<br />

to that odor than they were upon re-exposure to a different<br />

odor that had been paired with relaxation instructions. This apparent<br />

ability to induce stress below the level of awareness and its relevance to<br />

odor-conditioned stress responses as a potential mechanism underlying<br />

persistent aftereffects of trauma and odor-associated syndromes, such as<br />

multiple chemical sensitivity, will be discussed.<br />

Supported by DAMD17-01-2-0782<br />

118 Poster [ ] Clinical Evaluation and Consumer Research<br />

THE IMPACT OF MALINGERING ON THREE MEASURES OF<br />

OLFACTION<br />

Bailie J.M. 1, Rybalsky K. 1, Horning K.A. 1, Hoffman S.M. 1, Gesteland<br />

R.C. 2, Frank R.A. 1 1Psychology, University of Cincinnati, Cincinnati,<br />

OH; 2Cell Biology, University of Cincinnati, Cincinnati, OH<br />

Experts on the sense of smell are occasionally asked to evaluate<br />

olfaction in <strong>for</strong>ensic cases. Since it has been estimated that the faking of<br />

a physical injury, sensory or cognitive loss (known as malingering)<br />

occurs in as many as one out of six of these cases, it is valuable to have<br />

a measure of olfaction that is resistant to malingering. The Sniff<br />

Magnitude Test (SMT) is a newly-developed measure of olfactory<br />

function that uses sniffing responses to assess the sense of smell. The<br />

SMT may be resistant to malingering because it has low difficulty and<br />

at the same time it is not obvious to a typical patient how an olfactory<br />

loss is feigned. A total of 120 subjects were randomly assigned into<br />

four groups: a control group that received the standard administration of<br />

each olfactory test, Malingering Group 1 that received instruction to<br />

fake an olfactory loss but was not provided further instructions,<br />

Malingering Group 2 that was told to fake a loss and was given<br />

in<strong>for</strong>mation about what each test measures, and Malingering Group 3<br />

that was told to fake a loss and was coached on how to appear anosmic<br />

on each of the three tests. The SMT was compared to two traditional<br />

tests- PEA odor threshold and per<strong>for</strong>mance on the UPSIT. Results are<br />

discussed in terms of the impact of test in<strong>for</strong>mation on feigning an<br />

olfactory loss and the resistance of each test to malingering. This<br />

project supported by NIH SBIR grant DC04139, R. C. Gesteland, PI.<br />

30<br />

119 Poster [ ] Clinical Evaluation and Consumer Research<br />

UNILATERAL OLFACTORY THRESHOLDS IN A<br />

CHEMOSENSORY CLINICAL POPULATION<br />

Cowart B.J. 1, Pribitkin E. 2, Rosen D. 2, Klock C. 1, Laflam T. 1 1Monell<br />

Chemical Senses Center, Philadelphia, PA; 2Otolaryngology-Head &<br />

Neck Surgery, Thomas Jefferson University, Philadelphia, PA<br />

Although unilateral (UNI) testing of olfactory threshold sensitivity is<br />

routinely per<strong>for</strong>med in many chemosensory clinics to supplement<br />

bilateral (BI) tests, few studies have addressed the usefulness of these<br />

additional measures in either characterizing individual patients or<br />

providing insight into general characteristics of olfactory dysfunction.<br />

We report findings from 195 non-anosmic patients presenting to the<br />

Monell-Jefferson Taste & Smell Clinic who underwent both BI and<br />

UNI tests of thresholds <strong>for</strong> phenylethyl alcohol (PEA); BI and UNI<br />

thresholds <strong>for</strong> pyridine (PYR) were also obtained from a subset (n=73).<br />

37.4% of patients showed a left(L)-right(R) difference of at least one<br />

log step in PEA threshold concentration (v/v), but only 54.8% of those<br />

also showed a directionally consistent L-R difference in PYR<br />

thresholds. The presence of a UNI difference was not related to either<br />

degree of bilateral dysfunction or etiology; however, UNI testing did<br />

enable the identification of an olfactory problem in 15 patients whose<br />

BI testing yielded no evidence of abnormality. In contrast to what has<br />

been reported in largely non-clinical populations, patients´ best UNI<br />

thresholds <strong>for</strong> PEA were significantly poorer than their BI thresholds.<br />

Interestingly, however, this effect interacted with the presence of a L-R<br />

difference, being evident only in patients with comparable L-R<br />

thresholds. Thus bilateral facilitation of olfactory threshold sensitivity<br />

may occur in dysfunction when the two nostrils are similarly affected.<br />

Supported by NIH grant P50 DC00214.<br />

120 Poster [ ] Clinical Evaluation and Consumer Research<br />

CLINICAL EVALUATION OF THE SNIFF MAGNITUDE TEST<br />

Frank R.A. 1, Seiden A. 2, Bailie J. 1, Rybalsky K. 1, Gesteland R.C. 3<br />

1Psychology, University of Cincinnati, Cincinnati, OH; 2Otolaryngology<br />

and Maxillofacial Surgery, University of Cincinnati, Cincinnati, OH;<br />

3Cell Biology, Univ of Cincinnati, Cincinnati, OH<br />

The Sniff Magnitude Test (SMT) is a newly developed olfactory<br />

measure that examines sniff patterns to assess smell function. Its utility<br />

has been assessed in children, university students, and older adults<br />

where it has proven to be a reliable measure that is strongly correlated<br />

with the UPSIT and odor threshold tests. This project examined the use<br />

of the SMT in a clinical setting. 44 otolaryngology patients completed<br />

the UPSIT and SMT during a visit to an ENT clinic in Cincinnati, Ohio.<br />

Patients who came to the clinic <strong>for</strong> olfactory complaints were compared<br />

to patients that came in <strong>for</strong> sinus-related problems and patients with<br />

hearing-related complaints. It was predicted that the olfactory complaint<br />

group would per<strong>for</strong>m more poorly than the sinus and hearing groups on<br />

both the UPSIT and SMT. Patients with complaints about the sense of<br />

smell per<strong>for</strong>med significantly worse on both the UPSIT and SMT<br />

compared to the sinus patients [F(1, 30) = 9.89, p< .005 and F (1, 33) =<br />

9.90, p < .005, respectively]. This was also the case <strong>for</strong> the participants<br />

who had hearing complaints [F(1, 26) = 25.083, p < .001 and F(1, 29)<br />

=12.36, p < .001, respectively]. There was no difference between the<br />

sinus and hearing patients on the two olfactory tests. These results<br />

demonstrate the utility of the SMT in a clinical setting and support<br />

additional studies of its use in other clinical conditions where olfactory<br />

deficits are a concern. This project supported by NIH SBIR grant<br />

DC04139, R. C. Gesteland, PI.


121 Poster [ ] Clinical Evaluation and Consumer Research<br />

THE INFLUENCE OF ODOR PLEASANTNESS & IRRITATION<br />

ON THE SNIFF MAGNITUDE TEST<br />

Rybalsky K. 1, Bailie J. 1, Frank R.A. 1, Gesteland R.C. 2 1Psychology,<br />

University of Cincinnati, Cincinnati, OH; 2Cell Biology, Univ of<br />

Cincinnati, Cincinnati, OH<br />

The Sniff Magnitude Test (SMT) measures olfactory function based<br />

on the reduction of the size of a sniff in response to an odor. In the past,<br />

malodors have been successfully utilized to trigger sniff reduction.<br />

Experiments were conducted to assess the use of pleasant odors and to<br />

address concerns about possible irritation associated with the odorants<br />

being used with the SMT. 119 university students were tested in 3<br />

experiments. The stimuli included one malodor (methylthiobutyrate or<br />

MTB) and two pleasant odors, strawberry flavor (STR) and isoamyl<br />

acetate (AA). In Experiment 1, STR (1%v/v) suppressed sniffs by 37%.<br />

STR was then compared with MTB (3% v/v) and the results indicated<br />

that MTB suppressed sniffs by 46%, similar to but significantly more<br />

than STR (p < .05). In Experiment 2, assessment of sniff magnitude to<br />

MTB (1% v/v) and AA (1% v/v) yielded results similar to the STR.<br />

While sniff suppression to AA was effective (44%), participants<br />

suppress more to the MTB stimulus (51%, p < .05). Experiment 3<br />

assessed irritation and possible trigeminal sensitivity to MTB and AA.<br />

This was achieved using a nasal lateralization technique. It was<br />

demonstrated that neither MTB nor AA possess trigeminal components<br />

at the concentrations tested. The results from these studies support the<br />

use of both pleasant and unpleasant odors in the SMT and also allay<br />

fears about the potentially confounding role that trigeminal stimulation<br />

plays in the test. This project supported by NIH SBIR grant DC04139,<br />

R. C. Gesteland, PI.<br />

122 Slide [ ] Clinical Evaluation and Consumer Research<br />

INFLUENCES OF AGE AND SEX ON A<br />

MICROENCAPSULATED ODOR MEMORY TEST<br />

Choudhury E. 1, Moberg P. 2, Doty R. 3 1School of Dental Medicine,<br />

University of Pennsylvania, Philadelphia, PA; 2Department of<br />

Psychiatry, Neurology, and Otorhinolaryngology, University of<br />

Pennsylvania Medical Center, Philadelphia, PA; 3Smell and Taste<br />

Center, Department of Otorhinolaryngology, University of<br />

Pennsylvania Medical Center, Philadelphia, PA<br />

While influences of such variables as age and sex are well<br />

established <strong>for</strong> most standardized tests of odor identification and<br />

detection, this is not the case <strong>for</strong> tests of odor memory. In this study,<br />

231 non-smoking men and women, ranging in age from 10 to 68 years,<br />

were administered a standardized 12-item delayed match-to-sample<br />

micro-encapsulated odor memory test (OMT). Anosmics were excluded<br />

from the study. Each participant was asked to smell a target odor after<br />

its release from a micro-encapsulated odorant pad and then, after a<br />

delay interval of 10, 30, or 60 seconds, to pick the target from a<br />

similarly presented set of four odors, three of which were foils.<br />

Backward counting by threes was required during the delay intervals in<br />

an ef<strong>for</strong>t to minimize semantic rehearsal. Overall OMT scores were<br />

higher <strong>for</strong> women then <strong>for</strong> men, and decreased, in each sex, as a<br />

function of age in a manner similar to the age-related decline observed<br />

in tests of odor identification and detection. Per<strong>for</strong>mance did not change<br />

as a function of delay interval. A significant correlation between the<br />

overall OMT test scores and scores on the University of Pennsylvania<br />

Smell Identification Test were observed <strong>for</strong> women, but not <strong>for</strong> men, in<br />

accord with the notion that women may be more likely to employ<br />

semantic cues in their strategies to remember odors. The findings are<br />

discussed in light of the complexities of the construct of odor memory.<br />

This research was supported, in part, by the following grants from the<br />

National Institutes of Health, Bethesda, MD: PO1 DC 00161, RO1 DC<br />

04278, RO1 AG 17496 and RO1 MH 63381.<br />

31<br />

123 Poster [ ] Clinical Evaluation and Consumer Research<br />

USE, HANDLING AND ASSESSMENT OF PERFUMES<br />

DEPENDING ON BRAND NAME AND PACKING<br />

Buschmann-Maiworm R. 1, Henkel M. 2, Schurian W. 2 1Dept. of<br />

Psychology, University of Muenster, Münster, Germany; 2University of<br />

Muenster, Muenster, NRW, Germany<br />

The experiment investigates how the image of a perfume influences<br />

its assessment and handling.<br />

Three perfumes with different images, designed <strong>for</strong> different age<br />

groups, were tested. In a pilot study 130 subjects rated the general<br />

image of 10 different well known perfumes (e.g. age and style of users,<br />

wish to buy it).<br />

In the main experiment Chanel 5, Naomagic, Kölnisch Wasser were<br />

used. Each perfume was tested in its own bottle and packing and in the<br />

bottles of the other two. As a result we have 9 experimental conditions.<br />

225 women assessed the perfumes in 3 conditions in random order on<br />

bipolar rating scales while they were videotaped. The handling of the<br />

packing and the bottle were analysed (e.g. <strong>for</strong> duration of touching).<br />

Categorised <strong>for</strong> behavioural analysis were 4 ways of smelling the<br />

perfume: applying on a paper strip, applying on skin, smelling at the<br />

bottle, smelling at cap. Trials using Kölnisch Wasser were the shortest.<br />

The odor of the perfumes (Naomagic, Kölnisch Wasser and Chanel 5)<br />

in the bottle of Naomagic were much longer faned. The handling of the<br />

packing of Naomagic was longer than of Kölnisch Wasser. Kölnisch<br />

Wasser was rated as moderate pleasant. It was significantly rated as<br />

more pleasant in the packing of Naomagic and Chanel 5 than in its own<br />

package.<br />

124 Slide [ ] Odorant Receptors & Transduction<br />

ORGANIZATION OF CHEMOSENSORY SIGNALING<br />

COMPONENTS IN LIPID RAFTS<br />

Van Houten J. 1, Weeraratne S. 1, Chandran S. 1, Yano J. 1 1Biology,<br />

University of Vermont, Burlington, VT<br />

Glycosyl phosphatidylinositol (GPI) anchored proteins are tethered to<br />

the cell surface and function as receptors that generally activate tyrosine<br />

kinases to initiate signal transduction pathways in axon guidance or Tcell<br />

activation, among other functions. These proteins are characteristic<br />

of in lipid rafts, lipid domains that are enriched in glycosphingolipids<br />

and cholesterol and signal transduction proteins. At least one<br />

Paramecium receptor <strong>for</strong> folate is GPI anchored and likely to be in lipid<br />

rafts. Another component of the signal transduction pathway, the<br />

plasma membrane calcium pump (PMCA), is enriched in lipid rafts in<br />

other systems, and shows the hallmarks of lipid raft association in<br />

Paramecium. Fluorescence microscopy shows that the GPI anchored<br />

proteins, including the folate receptor, co-localize with PMCAs and<br />

basal bodies, consistent with the localization of these proteins at the<br />

bases of cilia. We are using cholera toxin to localize the gangliosides in<br />

rafts. In a separate approach to lipid rafts, we use gradients to separate<br />

membrane fractions by density, with lipid rafts and their proteins<br />

characteristically populating the lighter fractions. Raft fractions are<br />

those with gangliosides, cholesterol and GPI anchored proteins. There<br />

are light density fractions of the Paramecium membranes that include G<br />

proteins, the folate receptor, other GPI anchored proteins, the PMCAs.<br />

In a third approach, we remove cholesterol from the whole cell<br />

membranes and note that the cells are defective in their chemo-response<br />

to folate. Varied approaches help us to examine the role of lipid rafts<br />

and other Triton X-100 insoluble components, such as the cytoskeleton,<br />

in chemoresponse. GM 59988, DC 00721 and the Vermont Cancer<br />

Center.


125 Slide [ ] Odorant Receptors & Transduction<br />

DNA-BASED FLUORESCENT CHEMOSENSORS FOR DIRECT<br />

DETECTION OF VOLATILE COMPOUNDS IN AN<br />

ARTIFICIAL NOSE<br />

White J.E. 1, Williams L.B. 1, Atkisson M.S. 2, Kauer J.S. 1 1Neuroscience,<br />

Tufts University School of Medicine, Boston, MA; 2CogniScent, inc.,<br />

Weston, MA<br />

As reported previously (White et al., 2002, AChemS XXIV), we are<br />

developing a portable artificial olfactory system to detect and<br />

identify volatile compounds in the ambient environment. In a manner<br />

based directly on animal behavior, the device actively samples air via<br />

pulsatile sniffing. Air samples are drawn over an array of<br />

broadly-responsive, optically-based chemical sensors, which are used<br />

to detect and discriminate odorants in the sample. Device function<br />

thus depends directly on the sensitivity and diversity of the sensors<br />

in the array. Biopolymers, particularly nucleic acids, are attractive<br />

<strong>for</strong> building sensors of great variety and wide utility. This is<br />

because of the tremendous combinatorial complexity made possible<br />

by constructing them from sequences of just a few building blocks. We<br />

have found that dye-labeled DNA sensors, dried onto a substrate,<br />

differentially respond to volatile chemical compounds. To our<br />

knowledge, this is the first time that solid phase DNA-based sensors<br />

have been used <strong>for</strong> directly detecting small, vapor phase molecules.<br />

Furthermore, our experiments indicate that Cy3-labeled, single- strand<br />

DNA sequences of similar length but different base sequence can<br />

respond differently to the same odorant set. These observations<br />

suggests that "sensor discovery" via high throughput screening of<br />

large-scale DNA-based sensor libraries may be possible, thereby<br />

providing a rapid means to easily identify sensors that are optimized<br />

<strong>for</strong> defined odorant detection tasks. Supported by NIDCD and NSF.<br />

32<br />

126 Slide [ ] Odorant Receptors & Transduction<br />

MAKING SENSE OF OLFACTION THROUGH MOLECULAR<br />

MODELING<br />

Floriano W.B. 1, Hall S. 1, Leonard O. 2, Hummel P.A. 1, Vaidehi N. 1,<br />

Goddard W.A. 1 1Materials and Process Simulation Center, Cali<strong>for</strong>nia<br />

Institute of Technology, Pasadena, CA; 2University of Cali<strong>for</strong>nia -<br />

Berkeley, Berkeley, CA<br />

We used the MembStruk first principles computational technique to<br />

predict the three-dimensional (3D) structure of ten mouse olfactory<br />

receptors (S6, S18, S19, S25, S46, S50, S1, MOR-EV, MOR-EG, and<br />

MI7) <strong>for</strong> which experimental odorant recognition profiles are available.<br />

We used the HierDock method to scan each predicted OR structure <strong>for</strong><br />

potential odorant binding site(s), and to calculate binding energies of<br />

each odorant in these binding sites. The calculated binding affinity<br />

profiles correctly identify the chemical classes recognized<br />

experimentally by each OR, validating the predicted 3D structures and<br />

binding sites. Correlation between calculated binding affinity and<br />

elicited response is also found within each chemical class. However the<br />

cutoff response/no-response is not always well defined.<br />

For each of the ten ORs, the binding site is located between TMs 3<br />

through 6, with contributions from EC loops 2 and 3. In particular, we<br />

find six residue positions in TM3 and TM6 to be consistently involved<br />

in odorant binding. These positions are consistent with mutation data on<br />

ligand binding <strong>for</strong> other GPCRs and sequence hypervariability studies<br />

<strong>for</strong> ORs.<br />

Amino acid patterns associated with the recognition of chemical<br />

classes were defined using the predicted binding modes. These<br />

sequence fingerprints were used to probe the alignment of 869 OR<br />

sequences from the mouse genome in order to identify other ORs<br />

matching each fingerprint.<br />

This research was initiated with support by an ARO-MURI grant<br />

(Dr. Robert Campbell) and completed with finding from NIHBRGRO1-<br />

GM625523, NIH-R29AI40567, and NIH-HD36385. The computational<br />

facilities were provided by a SUR grant from IBM and a DURIP grant<br />

from ARO.<br />

127 Slide [ ] Odorant Receptors & Transduction<br />

THE MULTIFACETED RECEPTORS OF THE HUMAN NOSE<br />

Lancet D. 1, Alony R. 1, Atarot T. 1, Ben-Asher E. 1, Feldmesser E. 1, Gilad<br />

Y. 1, Khen M. 1, Man O. 1, Menashe I. 1, Olender T. 1, Stern S. 1 1Molecular<br />

Genetics, Weizmann Institute, Rehovot, Israel<br />

Human olfactory receptor (OR) genes are highly multifaceted, as<br />

manifested at several different levels. This is documented in the<br />

updated Human Olfactory Receptor Data Exploratorium (HORDE<br />

version 40, http://bip.weizmann.ac.il/HORDE/), which contains 853<br />

entries. One dimension of diversity spans a continuum between intact<br />

ORs and definite OR pseudogenes. To probe the undecided cases, we<br />

have devised methods to identify subtle deviations from shared motifs.<br />

Positional departures of initiation and stop codons may indicate inactive<br />

ORs. Second generation rhodopsin-based homology models provide<br />

new clues to sequence positions essential <strong>for</strong> functionally intact<br />

structure, including helix-kinking residues. A new algorithm, based on<br />

paralog-orthologs comparisons <strong>for</strong> three mammalian species (human,<br />

mouse and dog), allowed us to identify potential odorant contact<br />

residues. These harbor maximal intra-species variability, but also<br />

constitute additional targets <strong>for</strong> inactivating mutations. Another<br />

dimension of diversity is inter-individual variability, generated by<br />

single nucleotide polymorphisms that occur at pseudogenizing sequence<br />

positions. This results in a personal genetic “bar-code”, whereby every<br />

human individual has a different combination of intact and inactive<br />

ORs. These “segregating pseudogenes” are likely key determinants of<br />

specific anosmia. Finally, OR gene diversity is probed at the 5´<br />

untranslated exon sequences by computer-based and experimental<br />

approaches. This may shed light on presumed roles of upstream<br />

segments in OR messenger RNA stability and translation.


128 Slide [ ] Odorant Receptors & Transduction<br />

MOLECULES THAT REGULATE TRANSLOCATION AND<br />

FUNCTION OF MAMMALIAN ODORANT RECEPTORS<br />

Saito H. 1, Matsunami M. 1, Roberts R. 1, Chi Q. 1, Matsunami H. 1 1MGM,<br />

Duke University, Durham, NC<br />

Progress in understanding olfactory coding requires knowing ligand<br />

specificity of odorant receptors (ORs). It has been difficult to examine<br />

ligand-binding specificity of mammalian ORs in heterologous cells,<br />

mainly due to poor plasma membrane trafficking of ORs in these cells.<br />

Using a set of cDNA clones that are expressed by olfactory neurons, we<br />

have screened and found three genes that induce cell surface expression<br />

of ORs when expressed in HEK293T cells. We named them REEP1,<br />

<strong>for</strong> Receptor Expression Enhancing Protein 1, RTP1 and 2, <strong>for</strong><br />

Receptor Transporting Protein 1 and 2, respectively. All three proteins<br />

contain a single putative transmembrane domain and they are<br />

specifically expressed by the olfactory neurons in the olfactory<br />

epithelium. These proteins are associated with OR proteins, and<br />

enhance activation of OR by odorants. Using a cell line expressing<br />

these genes, we have identified new mouse ORs responding to aliphatic<br />

odorants. Our results suggest that REEP1, RTP1 and 2 have important<br />

roles in trafficking of ORs and provide an efficient system to<br />

investigate OR-odorant interaction.<br />

129 Slide [ ] Odorant Receptors & Transduction<br />

THE ROLE OF ODOR RECEPTORS IN ODOR CODING<br />

Hallem E.A. 1, Ho M.G. 1, Carlson J.R. 1 1MCDB, Yale University, New<br />

Haven, CT<br />

We are undertaking a large-scale analysis of Drosophila odor<br />

receptors. The Or gene family contains ~60 members, which are<br />

expressed in at least 35 functional classes of olfactory receptor neurons<br />

(ORNs). We are examining the functions of odor receptors using an<br />

“empty” ab3A antennal ORN (∆ab3A) that lacks its endogenous odor<br />

receptors as an in vivo expression system. Individual receptors are<br />

introduced into the empty ORN (∆ab3A:OrX) and odor response is<br />

assayed electrophysiologically. We are establishing a receptor-toneuron<br />

map by matching the odor response spectra of ∆ab3A:OrX<br />

ORNs to the odor response spectra of wild-type ORNs. We have found<br />

that the odor receptor dictates not only the odor response spectrum, but<br />

also the signaling mode (excitation v. inhibition) and the response<br />

dynamics of the ORN in which it is expressed. Different receptors,<br />

when expressed in the same ORN and given the same odorant stimulus,<br />

can confer responses that differ in signaling mode. Moreover, an<br />

individual receptor can confer responses of different modes to different<br />

odorants in the same ORN. The results thus show that odor receptors<br />

contribute to multiple aspects of odor coding in Drosophila ORNs, and<br />

they suggest a model <strong>for</strong> odor receptor transduction. Not only can<br />

multiple receptors function in an individual ORN, but an individual<br />

receptor can function in multiple ORNs, suggesting a broad<br />

compatibility among receptors and ORNs. Finally, we express two<br />

odor receptors from the malaria vector mosquito Anopheles gambiae in<br />

Drosophila and find that the female-specific receptor AgOr1 responds<br />

to 4-methylphenol, a component of human sweat.<br />

33<br />

Supported by NIH DC04729 and an NSF graduate fellowship.<br />

130 Slide [ ] Odorant Receptors & Transduction<br />

THE HOR17-4 SIGNALING SYSTEM – ONE RECEPTOR,<br />

DUAL CAPACITY<br />

Spehr M. 1, Schwane K. 1, Barbour J. 1, Riffell J.A. 2, Heilmann S. 3,<br />

Gisselmann G. 1, Hummel T. 3, Zimmer R.K. 2, Neuhaus E.M. 1, Hatt H. 1<br />

1Ruhr-Universität Bochum, Bochum, Germany; 2biology, University of<br />

Cali<strong>for</strong>nia, Los Angeles, CA; 3University of Dresden, Dresden,<br />

Germany<br />

A repertoire of mammalian olfactory receptor (OR) genes are<br />

predominantly expressed in spermatozoa. One of these receptors,<br />

named hOR17-4, is attributed an important role in human sperm<br />

physiology by mediating directed chemotactic movement in vitro.<br />

The underlying molecular mechanisms transducing OR ligand<br />

binding into flagellar motion remain obscure. Here, we show that sperm<br />

OR activation is coupled to a membrane-bound adenylyl cyclase<br />

(mAC). Odorant-induced Ca2+ signaling, chemotaxis, and<br />

chemokinesis are completely abolished by a specific inhibitor of mACs.<br />

Identifying membrane proteins in human sperm via mass spectometry,<br />

we show expression of specific receptors, G-proteins, and mACs. Their<br />

spatial distribution patterns largely correspond to the spatiotemporal<br />

character of odorant Ca2+ signals. Thus, our data link mAC activation<br />

to human sperm chemotaxis.<br />

Whether sperm ORs are functionally restricted to reproductive issues<br />

or additionally per<strong>for</strong>m their “conventional” task in olfaction is a<br />

longstanding question. There<strong>for</strong>e, we also investigate hOR17-4<br />

expression in human olfactory tissue adopting. Our results suggest that<br />

the same human OR protein is similarly utilized to fulfill chemosensory<br />

functions in such diverse cell types as spermatozoa and olfactory<br />

sensory neurons.<br />

131 Slide [ ] Odorant Receptors & Transduction<br />

MECHANISMS OF ODOR RECEPTOR GENE CHOICE<br />

Ray A. 1, Shiraiwa T. 1, Goldman A. 1, Van Der Goes Van Naters W. 1,<br />

Carlson J. 1 1MCDB, Yale University, New Haven, CT<br />

Little is known about how individual neurons select which Or genes<br />

to express. The Or gene family in Drosophila consists of ~60 members,<br />

most of which are expressed either in the antenna or the maxillary palp,<br />

but not both.<br />

We have used bioin<strong>for</strong>matics to identify cis -regulatory elements<br />

important <strong>for</strong> Or gene choice in the maxillary palp. Among several<br />

motifs that are over-represented upstream of Or genes, one,<br />

CTA(N) 9 TAA, is found within 500 bp upstream of Or genes expressed<br />

in the maxillary palp. We have found by mutational analysis that this<br />

motif is necessary <strong>for</strong> expression of Or genes in this organ. A second<br />

over-represented motif among maxillary palp genes is CTTATAA, and<br />

we have found evidence that this motif is a negative regulatory element<br />

that mediates repression of palp Or genes in the antenna.<br />

What in<strong>for</strong>mation specifies the particular neuron class in which a<br />

receptor is expressed? A pair of Or genes, Or85e and Or33c, is coexpressed<br />

in the pb2A neurons of the palp. We have identified 3 motifs<br />

shared between the upstream sequences of these two co-regulated<br />

genes, but not by other palp genes. To identify neuron-specific<br />

regulatory motifs <strong>for</strong> the remaining palp Or genes, we have compared<br />

the Or upstream regions with those of their D. pseudoobscura<br />

orthologs. This approach has identified several additional gene-specific<br />

conserved elements.<br />

A second mechanism is used to achieve co-regulation of two genes in<br />

another neuron, pb2B. Two clustered Or genes, Or46a and Or46b, are<br />

transcribed as a bicistronic message in this neuron. We are testing the<br />

activity of a putative IRES, and the functional significance of the co-


expression.<br />

132 Slide [ ] Odorant Receptors & Transduction<br />

REGULATION OF ODORANT RECEPTOR EXPRESSION<br />

Reed R.R. 1, Lewcock J.W. 2 1Molecular Biology and Genetics,<br />

HHMI/Johns Hopkins University, Baltimore, MD; 2Molecular Biology<br />

and Genetics, Johns Hopkins University, Baltimore, MD<br />

The expression of a single odorant receptor (OR) in each olfactory<br />

neuron from a repertoire of more than 1000 genes is essential <strong>for</strong> odor<br />

coding and axonal targeting. The genes encoding these receptors each<br />

possess a simple genomic structure and in several cases, small DNA<br />

segments surrounding the transcription initiation sites are sufficient to<br />

direct expression of reporters in a pattern that mimics the endogenous<br />

genes. One remarkable aspect of this gene regulation is that each<br />

olfactory neuron expresses OR protein from only one allele of this<br />

large, dispersed gene family. We have used targeted transgenesis,<br />

insertion of defined DNA constructs at specific location in the genome<br />

to identify a new role <strong>for</strong> OR protein as an essential regulator in the<br />

establishment of mono-allelic OR expression. OR-promoter driven<br />

reporters expresses in a receptor-like pattern, but unlike a native OR,<br />

are co-expressed with an additional OR allele. Expression of a<br />

functional OR from the identical promoter eliminates expression of<br />

other OR alleles. The presence of an untranslatable OR coding<br />

sequence in the mRNA is insufficient to exclude expression of a second<br />

OR. Together, these data identify the OR protein as a critical element in<br />

a feedback pathway that regulates odorant receptor selection. Current<br />

ef<strong>for</strong>ts in the laboratory are focused on elucidating the nature of the<br />

feedback signal and the mechanisms that lead to selective receptor<br />

expression.<br />

Supported by grants from NIDCD to R. R.<br />

133 Symposium [ ] The Ins and Outs of Sensory Cilia<br />

WHAT THE CHLAMYDOMONAS FLAGELLUM IS<br />

TEACHING US ABOUT SENSORY CILIA<br />

Witman G.B. 1 1Cell Biology, University of Massachusetts Medical<br />

School (Worcester), Worcester, MA<br />

Both motile cilia and non-motile cilia, including mammalian primary<br />

cilia, are sensory organelles that display receptors and relay signals<br />

about the extracellular environment to the cell body. The unicellular,<br />

biflagellate green alga Chlamydomonas reinhardtii has provided an<br />

essential foundation <strong>for</strong> understanding this important ciliary function.<br />

A notable example is the recent discovery and characterization of<br />

intraflagellar transport (IFT) in Chlamydomonas. IFT is a process in<br />

which flagellar precursors are actively moved into the flagellum and out<br />

to its tip, where axonemal assembly occurs; disruption of this process<br />

blocks flagellar assembly. Genetic and biochemical studies in<br />

Chlamydomonas have elucidated the molecular motors and other<br />

components of the IFT system; all of these proteins have homologues in<br />

other ciliated organisms, including C. elegans, D. melanogaster, and<br />

mammals. Because IFT is necessary <strong>for</strong> ciliary assembly, mutation of a<br />

gene encoding a mouse homologue of a Chlamydomonas IFT protein<br />

disrupts assembly of primary cilia, providing a valuable tool <strong>for</strong> testing<br />

the function of these widespread organelles (see abstract by G. Pazour).<br />

Although primary cilia cannot be isolated, Chlamydomonas flagella can<br />

be readily purified in amounts sufficient <strong>for</strong> biochemical analysis. An<br />

ongoing proteomic analysis of the Chlamydomonas flagellum is<br />

revealing numerous conserved proteins that are likely to be involved in<br />

sensory processes. The mammalian homologues of these proteins are<br />

prime candidates <strong>for</strong> carrying out sensory reception and signal<br />

transduction in the primary cilium. Supported by NIH GM 30626.<br />

34<br />

134 Symposium [ ] The Ins and Outs of Sensory Cilia<br />

PROBING THE FUNCTION OF MAMMALIAN PRIMARY<br />

CILIA BY ANALYSIS OF THE TG737 MOUSE<br />

Pazour G.J. 1 1Molecular Medicine, University of Massachusetts<br />

Medical School (Worcester), Worcester, MA<br />

Most mammalian cells have a non-motile primary cilium projecting<br />

from their surface. The importance of these organelles to mammalian<br />

health and development has been highlighted by recent analysis of the<br />

Tg737 mouse. This mouse has a mutation in the gene encoding the<br />

IFT88 subunit of the intraflagellar transport particle (Pazour et al., J.<br />

Cell Biol. 151:709-718) and develops polycystic kidney disease (PKD)<br />

(Moyer et al., Science 262:1329-1333) and other disorders. The Tg737<br />

mutation causes ciliary assembly defects in the kidney and other organs.<br />

These ciliary defects are likely to be the primary cause underlying the<br />

PKD and other diseases seen in the animal.<br />

We hypothesize that primary cilia are serving as organizing centers<br />

<strong>for</strong> sensory receptors and signaling proteins. In support of this, we and<br />

others have shown that the polycystins are localized on kidney primary<br />

cilia. The polycystins are membrane proteins thought to sense the state<br />

of the kidney epithelium and control proliferation and differentiation of<br />

these cells. Polycystin mutations cause excess cell proliferation in<br />

kidney nephrons resulting in adult onset PKD in humans. The ciliary<br />

assembly defect probably causes PKD by disrupting the localization of<br />

the polycystins.<br />

It is likely that all primary cilia serve similar functions in organizing<br />

receptors and pathways to monitor extracellular parameters that are<br />

important to the cell´s physiology.<br />

Work in my laboratory is funded by the NIH (GM60992) and the<br />

Worcester Foundation <strong>for</strong> Biomedical Research.<br />

135 Symposium [ ] The Ins and Outs of Sensory Cilia<br />

INTRAFLAGELLAR TRANSPORT MOTORS<br />

Scholey J.M. 1, Ou G. 1, Snow J. 1, Gunnarson A. 1 1Center <strong>for</strong> Genetics<br />

and Development and Section of Molecular and Cellular Biology,<br />

University of Cali<strong>for</strong>nia, Davis, Davis, CA<br />

The assembly and function of sensory cilia depends upon<br />

intraflagellar transport (IFT), the bidirectional transport of<br />

macromolecular complexes called IFT-particles, along axonemal<br />

microtubules (MTs) between the base of the cilium and the distal tip.<br />

We are studying the role of IFT in the <strong>for</strong>mation and function of the<br />

sensory cilia on the endings of chemosensory neurons within the<br />

nematode, C. elegans. These cilia serve as specialized compartments <strong>for</strong><br />

concentrating the sensory signaling machinery that detects chemical<br />

cues in the environment and thereby play important roles in<br />

chemosensory behaviour. Our hypothesis is that IFT-particles<br />

contribute to ciliary function by carrying key components of the ciliary<br />

axoneme, the signaling machinery, and possibly signals themselves,<br />

between the base and the tip of the cilium, and that the transport of<br />

these particles depends upon the action of two anterograde motors,<br />

kinesin-II and OSM-3-kinesin and a retrograde motor, IFT-dynein. We<br />

will report our recent progress in using light microscopy, biochemistry,<br />

genomics and genetics to dissect the protein machinery that drives IFT<br />

in this system.<br />

Reference: Scholey JM, 2003, Intraflagellar Transport. Ann Rev.<br />

Cell Dev. Biol., 19, 423.


136 Poster [ ] Olfactory Development<br />

METAMORPHOSIS OF AN OLFACTORY SYSTEM:<br />

HORMONAL REGULATION OF GROWTH AND<br />

PATTERNING IN THE ANTENNAL IMAGINAL DISC OF THE<br />

MOTH MANDUCA SEXTA.<br />

Fernandez K. 1, Bohbot J. 1, Vogt R. 1 1Biological <strong>Sciences</strong>, University of<br />

South Carolina, Columbia, SC<br />

Peripheral olfactory systems of insects such as moths, bees,<br />

mosquitoes and flies undergo metamorphosis, trans<strong>for</strong>ming from a<br />

simple larval antenna to the highly complex adult antenna mediating<br />

diverse chemosensory behaviors. Adult antennae derive from imaginal<br />

discs which grow during the larval stage, and undergo neurogenesis and<br />

morphogenesis during the pupal stage. We are characterizing patterns<br />

of morphogenic and gene expression activities in the imaginal disc and<br />

early developing antenna to identify events which lead to the patterning<br />

of the adult antenna and the specific expression of genes (ORs, OBPs)<br />

by distinct classes of olfactory sensilla. This study focuses on the<br />

antennal disc; disc growth occurs during the final larval instar from a<br />

ring of tissue surrounding the base of the larval antenna. We have<br />

characterized the spatial patterns of disc growth using an antibody<br />

against phosphorylated histone H3 (mitotic marker) and have<br />

characterized temporal and spatial specific expression of transcription<br />

factors Broad (metamorphosis) and distal-less (pattern) and the<br />

signaling protein Notch. Prior to pupation, the disc elongates and<br />

everts; we have shown this process is regulated by ecdysteroid<br />

hormones, and have observed a subsequent decline in total DNA<br />

content suggesting apoptotic events associated with this restructuring.<br />

These studies are establishing a foundation <strong>for</strong> identifying early events<br />

leading to the selection of specific chemosensory phenotypes of adult<br />

olfactory sensilla.<br />

137 Poster [ ] Olfactory Development<br />

IG-FAMILY CELL ADHESION MOLECULES IN THE<br />

DEVELOPING ANTENNAL LOBE OF THE MOTH MANDUCA<br />

SEXTA<br />

Gibson N.J. 1, Tolbert L.P. 1 1ARL Division of Neurobiology, University<br />

of Arizona, Tucson, AZ<br />

Establishment of the adult antennal (olfactory) lobe of Manduca<br />

sexta requires numerous choices regarding pathfinding, migration,<br />

fasciculation, and branching by olfactory receptor cells, glia, and<br />

antennal lobe neurons. We are investigating the roles played by cell<br />

adhesion molecules in this process. Using an antibody to Manduca<br />

neuroglian, a homolog of vertebrate L1, we find that receptor cell axons<br />

express neuroglian specifically in the axonal sorting zone and in the<br />

nerve layer surrounding the glomerular layer in the lobe; expression is<br />

notably absent in the glomerular neuropil. Neuropil-associated glia also<br />

label, and do so even in the absence of afferent axons. This suggests the<br />

possibility of a homophilic interaction between neuroglian molecules<br />

on receptor cell axons and antennal-lobe glia. Following sorting of<br />

axons into new fascicles, we find that only some fascicles are labeled,<br />

suggesting targeting-related specificity, as has been seen <strong>for</strong> fasciclin II.<br />

In a search <strong>for</strong> additional cell adhesion molecules, we have used an<br />

antibody against the Ig-domain of human NCAM and obtained strong<br />

labeling of antennal lobe neurons and glia, but not axons. Western blots<br />

using this antibody demonstrate a prominent band at 125 KDa, similar<br />

in size to vertebrate NCAM-120 and OCAM and distinguishing this<br />

protein from the other Ig-containing molecules known to occur in<br />

Manduca, fasciclin II and neuroglian. Supported by NIH DC2004598.<br />

35<br />

138 Poster [ ] Olfactory Development<br />

OLFACTORY RECEPTOR CELLS OF TRANS-SEXUALLY<br />

GRAFTED FEMALE ANTENNAE DETERMINE ODOR<br />

RESPONSES OF OUTPUT NEURONS IN THE ANTENNAL<br />

LOBE OF MALE MANDUCA SEXTA<br />

Reisenman C.E. 1, Stein H. 1, Christensen T.A. 1, Hildebrand J.G. 1 1ARL<br />

Division of Neurobiology, University of Arizona, Tucson, AZ<br />

The antennal lobe (AL) of the moth Manduca sexta contains a small<br />

number of sexually dimorphic glomeruli: the macroglomerular complex<br />

(MGC) in males and the large female glomeruli (LFGs) in females. The<br />

role of olfactory receptor cells (ORCs) in the <strong>for</strong>mation of these<br />

glomeruli was demonstrated by trans-sexual transplantation<br />

experiments: sensory axons of a grafted male antenna induce <strong>for</strong>mation<br />

of an MGC in a host female, while sensory axons of a grafted female<br />

antenna induce <strong>for</strong>mation of LFGs in a host male. The odor-response<br />

properties of the induced glomeruli, however, are unknown. Using<br />

intracellular recording and staining techniques, we studied the<br />

properties of output neurons (PNs) in males with a grafted female<br />

antenna. To produce these gynandromorphs, one antennal disk from a<br />

2 nd -day 5 th -instar larva was excised and replaced by the corresponding<br />

disk from a female donor. 22% of animals had a normal female antenna<br />

(n=7) and an antennal nerve innervating the AL. We found that PNs<br />

arborizing in the induced LFGs (like latLFG-PNs in normal females)<br />

were more sensitive to (+) than to (-)linalool (n=3, 2 animals).<br />

Although not morphologically identified, we found 3 more presumptive<br />

PNs with the same odor specificity. The morphology of PNs<br />

innervating glomeruli outside the induced LFGs was similar to that of<br />

PNs in normal females (n=4). One of these PNs arborizes in a<br />

glomerulus next to the induced latLFG, and did not respond to antennal<br />

stimulation with either (+) or (-)linalool. These preliminary results are<br />

consistent with the hypothesis that ORCs confer specific odor-tuning<br />

characteristics to their glomerular targets.<br />

Supported by NIH/PEW


139 Poster [ ] Olfactory Development<br />

NO-MEDIATED SIGNALING FROM OLFACTORY<br />

RECEPTORS TO PERIPHERAL NERVE GLIA IN THE MOTH<br />

OLFACTORY PATHWAY<br />

Oland L.A. 1, Gibson N.J. 2, Tolbert L.P. 2 1A.R.L. Division of<br />

Neurobiology, University of Arizona, Tucson, AZ; 2Neurobiology,<br />

University of Arizona, Tucson, AZ<br />

In the developing olfactory (antennal) nerve of the moth Manduca<br />

sexta, the glia that populate the nerve arise in the antenna and migrate<br />

along the nerve. They reach the CNS/PNS interface region of the nerve<br />

several stages after the first axons have arrived and begin to ensheathe<br />

bundles of axons in the nerve. In recent in vitro studies, antennal nerve<br />

(AN) glia were found to advance readily along glial processes, and<br />

ORN axons were found to induce the AN glia to <strong>for</strong>m multi-cellular<br />

arrays (Tucker and Tolbert, 2003). Array <strong>for</strong>mation did not require<br />

direct ORN-AN glia contact, but occurred only in glia in close<br />

proximity to the axons, suggesting that the axons release a short-range<br />

diffusible signal. Evidence from previous studies of nitric oxide (NO) in<br />

developing Manduca indicated that a Ca ++ -dependent NO synthase is<br />

strongly expressed in the antenna (Nighorn et al., 1998), that the ORN<br />

axons express NO synthase during stages of axon ingrowth when the<br />

AN glia are populating the nerve (Gibson and Nighorn, 2000), and that<br />

NO regulates AN glial migration (Gibson et al., 2001). Current<br />

experiments test whether the <strong>for</strong>mation of arrays by AN glia in vitro is a<br />

consequence of NO release by ORN axons. In co-cultures of ORN<br />

neurons and PN glial cells, glial arrays can by reduced by a NOS<br />

inhibitor. In cultures of AN glia only, NO donors can induce glial<br />

chaining. Our initial results suggest that NO signaling from the ORN<br />

axons may be involved in the normal developmental processes of AN<br />

glia elongation and migration. Supported by NIH-DC04598 to LPT.<br />

140 Poster [ ] Olfactory Development<br />

EXPRESSION PROFILES OF GENES REGULATED BY<br />

THYROID HORMONE IN THE NOSE OF XENOPUS LAEVIS<br />

Walworth E. 1, Burd G. 1 1Molecular and Cellular Biology, University of<br />

Arizona, Tucson, AZ<br />

In Xenopus, metamorphosis is initiated by rising levels of<br />

endogenous thyroid hormone (T ). During metamorphosis from tadpole<br />

3<br />

to frog, essentially every organ/tissue is remodeled and many new<br />

structures must be <strong>for</strong>med. The nose of Xenopus tadpoles contains two<br />

areas of sensory epithelium. At metamorphosis, one area is completely<br />

remodeled and a new, third area of sensory epithelium is <strong>for</strong>med de<br />

novo. Apoptosis, cell proliferation, remodeling of the extracellular<br />

matrix and cell differentiation all take place during this trans<strong>for</strong>mation.<br />

Previously by our laboratory, more than 125 gene fragments associated<br />

with metamorphosis in the frog nose were isolated and identified using<br />

representational difference analysis, including gene fragments known to<br />

be associated with all of the above cellular processes. In this study, 56<br />

of those genes were subjected to real time quantitative RT-PCR at 5<br />

different stages of development throughout metamorphosis. We found<br />

the expression levels of several of the T early response genes,<br />

3<br />

including Gene 12, transcription factors (nuclear factor I-B, TH/bZIP,<br />

TRβ) and an extracellular matrix remodeling enzyme (stromelysin 3),<br />

were elevated 3 to 57 fold in response to 48 hours of T treatment, and<br />

3<br />

expression remained elevated throughout metamorphic climax. The<br />

elevated levels of expression were visualized by in situ hybridization.<br />

There<strong>for</strong>e, these genes which are known to participate in remodeling<br />

other tissues undergoing metamorphosis, also aid in remodeling the<br />

nasal capsule. Support: NIDCD #DC03905<br />

36<br />

141 Poster [ ] Olfactory Development<br />

DE NOVO DNA METHYL TRANSFERASES AND METHYL<br />

DNA BINDING DOMAIN PROTEINS IN OLFACTORY<br />

NEUROGENESIS<br />

Macdonald J. 1, Gin C. 1, Roskams A.J. 1 1Zoology, University of British<br />

Columbia, Vancouver, British Columbia, Canada<br />

Epigenetic Regulation (Silencing) of mammalian genomic DNA<br />

occurs when de novo methyltransferases (DNMTs) methylate sites in<br />

the mammalian genome to <strong>for</strong>m targets <strong>for</strong> methyl-CpG binding<br />

domain proteins (MBDs). MBDs <strong>for</strong>m repressor complexes that can<br />

directly repress promoters or modify chromatin structure by recruitment<br />

of histone deacetylases (HDACs). In the mouse olfactory epithelium,<br />

DNMT1, 2, 3a, and 3b are expressed in the OE from embryonic<br />

development through adulthood. DNMT3a and 3b expression is<br />

maximal in developing ORNs at peaks of neurogenesis. DNMT3b is<br />

expressed in presumptive globose basal cells, sustentacular cells, and<br />

cells migrating from the basal layer to the sustentacular layer in the<br />

E16.5 OE. DNMT3b is primarily expressed (with PCNA) in actively<br />

proliferating cells and may thus play a role in ORN progenitor cell<br />

cycle exit or neuronal fate choice. DNMT3b+/PCNA+ cells migrating<br />

from the basal layer are NST negative and likely represent an<br />

alternative lineage to the neuronal lineage. DNMT3a is expressed in<br />

post-mitotic immature ORNs and a subset of mature ORNs and may<br />

play a role in lineage restriction and induction of terminal ORN<br />

differentiation in non-proliferating neuro-precursors. Expression of<br />

HDAC2 is induced at the earliest stages of neuronal differentiation and<br />

is down-regulated at terminal differentiation of ORNs. A small<br />

percentage of DNMT3b positive cells begin to express HDAC2, while<br />

the majority of DNMT3a positive cells express HDAC2. This<br />

continuum of HDAC2 expression suggests that it may be appropriately<br />

expressed to mediate gene silencing responses to DNA methylation<br />

catalyzed by both DNMT3b and DNMT3a.<br />

142 Poster [ ] Olfactory Development<br />

MECHANISMS BY WHICH BDNF AND NGF ACT AND<br />

INTERACT WITH ATRIAL NATRIURETIC PEPTIDE TYPE-C<br />

TO PROMOTE PROLIFERATION OR DIFFERENTIATION OF<br />

OLFACTORY NEURONAL PRECURSORS.<br />

Simpson P.J. 1, Moon C. 1, Ronnett G.V. 1 1Neuroscience, Johns Hopkins<br />

University, Baltimore, MD<br />

Both BDNF and NGF promote the proliferation of a subset of<br />

olfactory neuronal precursor cells. Concurrent exposure to atrial<br />

natriuretic peptide type-C (CNP) inhibits this proliferation and<br />

promotes differentiation. Immunocytochemical analysis of cell cycle<br />

regulatory proteins shows that both BDNF and NGF increase levels of<br />

cyclin D1 and cdk4 although the mechanisms behind these increases<br />

differ. NGF also increases p27. Addition of CNP with neurotrophin<br />

inhibits increases in cyclin D1 and p27 while alternatively increasing<br />

levels of a variety of cell cycle inhibitory proteins, including p21. The<br />

profile and time course of this inhibitor expression varies between<br />

BDNF and NGF. Analysis of the mechanisms behind changes in cyclin<br />

D1 and p21 shows that inhibition of the MAPK pathway blocks<br />

alterations in cyclin D1 levels in response to either neurotrophins or<br />

neurotrophins with CNP. CNP however does not promote Erk1/2<br />

phosphorylation, suggesting that it may alter the specificity of MEK1/2<br />

signaling. Other signal transduction pathways are responsible <strong>for</strong><br />

alterations in p21. Inclusion of cycloheximide results in an increase in<br />

p21 levels in response to neurotrophins and an increase in cyclin D1<br />

levels in response to neurotrophins with CNP. These results are<br />

mimicked by application of 26S proteosome inhibitors. This suggests<br />

that the switch between proliferation and differentiation is primarily<br />

regulated through reciprocal degradation of inhibitory or progressive<br />

cell cycle proteins with a requirement <strong>for</strong> protein synthesis. Supported<br />

by NIDCD 5RO3DC005704-02 (PJS).


143 Poster [ ] Olfactory Development<br />

ADULT OLFACTORY PROGENITOR CELLS GIVE RISE TO<br />

BOTH NEURONS AND NON-NEURONAL CELLS IN<br />

CULTURE<br />

Jang W. 1, Woo J. 1, Schwob J.E. 1 1Anatomy and Cellular Biology, Tufts<br />

University, Boston, MA<br />

Recent data, including transplantation of FACS-sorted cells, indicate<br />

that the adult olfactory epithelium (OE) retains mutipotent progenitors<br />

(MPPs) among the globose basal cell (GBC) population, which can be<br />

activated when reconstitution of the OE requires replacement of both<br />

neurons and non-neuronal cells. Not much is known about the<br />

regulation of progenitor cell capacity, and in vitro studies of<br />

multipotency are warranted. Dissociated epithelial cells were taken<br />

from mice exposed to MeBr gas (MeBr-OE) 40 hr prior to use.<br />

Harvested cells were plated either on uncoated or matrigel-coated<br />

surfaces and grown in complex, serum-rich medium or defined, serumfree<br />

medium. In serum-rich medium, regardless of coating, MeBr-OE<br />

gives rise to cultures that are heterogeneous: constituent cells, in<br />

aggregate, display the phenotypic characteristics of neurons (TuJ-1[+]),<br />

sustentacular cells (cytokeratin [CK]-18[+]), and horizontal basal cells<br />

(CK-5 or 14[+]). TuJ-1 (+) cells grow in a scattered fashion, while CKexpressing,<br />

epithelioid cells cluster together tightly. In serum-free<br />

medium, MeBr-OE generates only CK-18 or CK-5/14 (+) epithelioid<br />

islands. Our GBC markers GBC-2 and GBC-3 stain some, but not all,<br />

cells in the islands. Interestingly, cells grown on matrigel, regardless of<br />

serum, generate spheres that project above the surface to the dish. The<br />

ability to assay and manipulate adult MPPs in vitro will give us a better<br />

understanding of the mechanisms that regulate proliferation and<br />

differentiation, and may lead, eventually, to therapeutic applications.<br />

Supported by R01 DC02167 and R21 DC006517<br />

144 Slide [ ] Olfactory Development<br />

STUDIES OF OLFACTORY CELL LINEAGE AND<br />

DIFFERENTIATION USING AN IN VITRO NEUROSPHERE<br />

MODEL AND TIME-LAPSE VIDEOMICROSCOPY<br />

Cunningham A.M. 1, Marlicz W. 2 1Developmental Neurosciences<br />

Program, University of New South Wales, Sydney, NSW, Australia;<br />

2University of New South Wales, Sydney, NSW, Australia<br />

The mammalian olfactory neuroepithelium possesses a unique<br />

progenitor cell as it avidly supports neurogenesis throughout adulthood<br />

and is capable of reconstituting cells of neuronal and non-neuronal<br />

lineages after injury. Most evidence suggests this progenitor resides in<br />

the GBC layer and using retroviral lineage tracing Huard et al. (1998)<br />

found evidence supporting a multipotent progenitor capable of making<br />

GBCs, HBCs and sustentacular cells. To identify this cell in vitro we<br />

developed a culture system of olfactory progenitor cells. Turbinate<br />

tissue was taken from 48 hr old Wistar rat pups and putative progenitors<br />

enriched from a preparation of dissociated olfactory neurons<br />

(Cunningham et al. 1999). Clonal neurospheres <strong>for</strong>med from individual<br />

motile cells and made progeny of neuronal and sustentacular classes,<br />

based on immunocytochemical analysis <strong>for</strong> GBC-1, Olf-1, G , Type 3<br />

olf<br />

adenylyl cyclase and SUS-4. Double-labeling immunocytochemistry of<br />

young spheres revealed cellular heterogeneity, consistent with lineage<br />

specification occurring early in sphere development. Spheres passaged<br />

and generated secondary spheres although at lesser frequency than CNS<br />

neurospheres generated from <strong>for</strong>ebrain. Our data is consistent with our<br />

having isolated in vitro a progenitor predicted to exist by Huard et al.<br />

(1998). Our system of clonal neurospheres provides a model of<br />

progenitor cell development that will allow better understanding of the<br />

mechanisms underlying olfactory neurogenesis. Supported by the<br />

Garnett Passe and Rodney Williams Memorial Foundation<br />

37<br />

145 Slide [ ] Olfactory Development<br />

STOCHASTIC YET BIASED EXPRESSION OF MULTIPLE<br />

DSCAM SPLICE VARIANTS BY INDIVIDUAL CELLS<br />

Chess A. 1, Daly M. 2, Neves G. 2, Zucker J. 2 1Biology, Massachusetts<br />

Institute of Technology, Cambridge, MA; 2Whitehead Institute,<br />

Cambridge, MA<br />

The Drosophila Dscam gene is essential <strong>for</strong> axon guidance and has<br />

38,016 possible alternative splice <strong>for</strong>ms. This extraordinary diversity<br />

can potentially be used to distinguish cells. We have analyzed the<br />

Dscam mRNA iso<strong>for</strong>ms expressed by different cell types and individual<br />

cells. The choice of splice variants expressed is regulated both spatially<br />

and temporally. Different subtypes of photoreceptors express broad yet<br />

distinctive spectra of Dscam iso<strong>for</strong>ms. Single cell RT-PCR documented<br />

that individual cells express at least several different Dscam iso<strong>for</strong>ms<br />

and allowed an estimation of the diversity that is present. For example,<br />

we estimate that each R3/R4 photoreceptor cell expresses 14-50 distinct<br />

mRNAs chosen from the spectrum of thousands of splice variants<br />

distinctive of its cell type. Thus, every cell´s Dscam repertoire is<br />

different from those of its neighbors providing a potential mechanism<br />

<strong>for</strong> the generation of unique cell identity in the nervous system and<br />

elsewhere. Prior studies by Zipursky and colleagues indicate an<br />

important role <strong>for</strong> the Dscam gene in controlling the proper targeting of<br />

olfactory receptor neurons. We are currently assessing the importance<br />

of the expression of distinct alternative splice <strong>for</strong>ms by different<br />

olfactory neurons in axon targeting specificity.<br />

146 Poster [ ] Olfactory Development<br />

A PUTATIVE ROLE FOR MHCI IN THE AXONAL<br />

TARGETING OF THE MOUSE OLFACTORY SYSTEM<br />

Salcedo E. 1, Restrepo D. 1 1Cell and Developmental Biology, University<br />

of Colorado Health <strong>Sciences</strong> Center, Denver, CO<br />

The ability to smell offers a remarkable solution to the daunting<br />

challenge of discriminating between diverse chemical molecules.<br />

Investigation into the cellular and molecular mechanisms of<br />

mammalian olfaction has established the olfactory bulb as a primary<br />

processing unit that organizes the signaling from olfactory sensory<br />

neurons into a topographical sensory map. However, the molecular<br />

mechanisms establishing this sensory map remain to be elucidated. In<br />

the current study, we investigate the role that major histocompatibility<br />

complex class I molecules (MHCI) play in organizing this sensory map.<br />

Towards this end, we have used immunohistochemistry to demonstrate<br />

the expression of MHCI molecules in the main olfactory bulb (MOB).<br />

Additionally, we have examined mice that are severely deficient in the<br />

expression of MHCI on the surfaces of cells (TAP -/- ). By crossing these<br />

mice with a strain of mice that co-expresses Tau-LacZ with the P2<br />

olfactory receptor, we can visualize the axonal projection patterns of the<br />

P2 olfactory sensory neurons in a background deficient <strong>for</strong> the<br />

expression of MHCI molecules. After mapping the P2 glomeruli in the<br />

MOB of both wild-type and TAP -/- mice, we have uncovered subtle but<br />

statistically significant differences in the location and number of P2<br />

glomeruli.


147 Poster [ ] Olfactory Development<br />

CELL TYPES EXPRESSING OMP IN THE OLFACTORY<br />

EPITHELIUM OF LARVAL ZEBRAFISH<br />

Sakata Y. 1, Michel W.C. 1 1Physiology, University of Utah, Salt Lake<br />

City, UT<br />

The olfactory epithelium of zebrafish contained ciliated, microvillar<br />

and crypt-type sensory neurons. While there appears to be some overlap<br />

in specificity across the cell types the ciliated OSNs appear to respond<br />

preferentially to bile salts and the microvillar cells to amino acids.<br />

Recently, Celik et al (Euro. J. Neurosci. 15:798, 2002) reported the<br />

expression of GFP under the zebrafish OMP promoter was found<br />

primarily in ciliate OSNs but also found in cells with shorter and stouter<br />

dendrites, presumably microvillar cells. In the current investigation we<br />

produced a genetic construct with the zebrafish OMP promoter driving<br />

expression of eYFP to quantitatively examine the distribution of eYFP<br />

in the OSN populations of larval zebrafish. The zOMP promoter was<br />

cloned from genomic DNA using primers described by Celik et al<br />

(2002), ligated into an pEYFP vector (Clontech). The vector was<br />

subsequently linearized and injected into 1 cell stage zebrafish<br />

embryos. OMP promoter driven expression of eYFP was noted by 48<br />

hours. The transiently transfected embryos were reared <strong>for</strong> 48-72 hours<br />

post-fertilization, fixed and processed <strong>for</strong> whole embryo<br />

immunocytochemistry using the anti-GFP antibody. After<br />

immunostaining the embryos were embedded in Eponate plastic and<br />

sectioned <strong>for</strong> electron microscopy. Preliminary counts indicate that the<br />

majority of the eYFP expressing cells are ciliated OSNs but confirm the<br />

expression was also observed in some microvillar OSNs. Until a stable<br />

transgenic line is produced we cannot provide reliable estimates of the<br />

proportions of ciliated or microvillar OSNs expressing. Supported by<br />

NIH DC01418 and NS-07938.<br />

148 Poster [ ] Olfactory Development<br />

TYROSINE HYDROXYLASE-LIKE IMMUNOREACTIVE<br />

CELLS IN THE OLFACTORY TRACTS OF GOLDFISH<br />

Hansen A. 1, Finger T.E. 2 1University of Colorado Health <strong>Sciences</strong><br />

Center, Denver, CO; 2Cell and Developmental Biology, University of<br />

Colorado Health <strong>Sciences</strong> Center, Denver, CO<br />

Goldfish possess a substantial population of tyrosine hydroxylase<br />

(TH)-like-immunoreactive cells within the olfactory tracts (OT) as well<br />

as TH-like-ir cells in the olfactory bulbs (OB). The TH-like-ir neurons<br />

in the OB have round cell bodies and large round nuclei with a single<br />

dendritic process extending towards the margin of the OB. These<br />

neurons were identified as a subset of granule cells (Alonso et al.,<br />

1989). The nature and origin of the TH-like-ir cells in the OT are<br />

unknown. The aim of the present study was to clarify whether the THlike-ir<br />

cells are newly generated cells which migrate into the OB, or<br />

whether they are more mature cells that are stationary in the OT.<br />

Goldfish was chosen as a model since they have long OTs, and ample<br />

in<strong>for</strong>mation is available as to anatomy, physiology, and behavior.<br />

BrdU-injections were used to visualize newly generated cells; and<br />

antibodies used to characterize the nature of the cells in question. The<br />

TH-like-ir cells in the OT are similar in shape to those in the OB,<br />

however, the tract cells are bipolar, extending one long process towards<br />

the OB and one towards the telencephalon. The majority of these cells<br />

occur in the medial OT, fewer in the lateral OT. One day after injection,<br />

BrdU-positive nuclei occur along the medial and lateral OTs. Six days<br />

after BrdU-injection, labeled nuclei are present at the whole length of<br />

the OTs and also in the OB. Further experiments are underway to test<br />

whether the TH-positive cells of the OT are equivalent to the rostral<br />

migratory stream of rodents.<br />

This study was supported by NIDCD grant RO1 DC033792 to J.<br />

Caprio and P30 DC04657 to D. Restrepo.<br />

38<br />

149 Slide [ ] Olfactory Development<br />

POSTNATAL CHANGES IN THE RAT MODIFIED<br />

GLOMERULAR COMPLEX: A QUANTITATIVE<br />

CYTOCHROME OXIDASE STUDY<br />

Meisami E. 1 1Molecular and Integrative Physiology, University of<br />

Illinois at Urbana-Champaign, Urbana, IL<br />

The modified glomerular complex (MGC) has been described as a<br />

set of glomeruli on the dorsomedial side of the rat olfactory bulb (OB),<br />

associated with suckling behavior in the neonate. We further explored<br />

the morphometric and anatomical features of MGC during postnatal<br />

development, using coronal series of cytochrome oxidase (CO) stained<br />

sections of the OB at postnatal ages of 1-, 3 -, 10- and 25 -days and<br />

analyzed with regard to lateral and medial distribution of the glomeruli,<br />

their size and number, and the anterior-posterior (AP) length of MGC.<br />

The MGC stained intensely <strong>for</strong> CO even in the newborn animal,<br />

compared to main OB glomeruli. In addition to the originally described<br />

medial complex, a lateral complex was also noted, particularly in rat<br />

pups 3 days and older. This complex had fewer glomeruli and was<br />

shorter than the medial complex, but its glomeruli were similar in size<br />

and stain intensity. Anteriorly the lateral and medial portions appear to<br />

fuse along the midline, <strong>for</strong>ming an asymmetrical horseshoe structure.<br />

Total glomeruli number in the entire complex (medial and lateral)<br />

increased postnatally, from about 15 at birth to about 20 at 3 days, 40 at<br />

10 days, and about 70 at day 25. During the same period, mean<br />

glomerular diameter increased from about 60 mm at birth to about 85<br />

mm in the weanling. These changes were accompanied by marked<br />

increases in total glomerular volume. Results indicate that MGC is well<br />

developed at birth, but continues to develop postnatally in terms of<br />

number and size of its glomeruli, indicating possible olfactory functions<br />

of this complex in the weanling and older animals.<br />

Support: University of Illinois Research Funds<br />

150 Poster [ ] Odor Activation and Modulation<br />

OMP IS A MODULATOR OF CA2+ CLEARANCE PROCESSES<br />

IN MOUSE OLFACTORY RECEPTOR NEURONS (ORNS)<br />

Kwon H.J. 1, Leinders-Zufall T. 1, Zufall F. 1, Margolis F.L. 1 1Dept.<br />

Anatomy & Neurobiology, Program in Neuroscience, University of<br />

Maryland, Baltimore, MD<br />

Ca 2+ entry to ORNs is a well-studied event in chemosensory<br />

transduction. However, restoration of intracellular Ca 2+ levels to the<br />

pre-stimulus level is less well understood. Recent behavioral and<br />

electrophysiological analyses of OMP–null mice led us to hypothesize<br />

that OMP participates in the recovery phase of olfactory signaltransduction.<br />

To investigate this we compared Ca 2+ transients in ORNs<br />

of wild-type and OMP-null mice by activating various pathways to<br />

increase the cytoplasmic [Ca 2+ ]. KCl, 3-isobutyl-1-methylxanthine<br />

(IBMX), and caffeine were used to induce Ca 2+ influx from voltagegated<br />

Ca 2+ channels, cyclic-nucleotide gated channels, and intracellular<br />

Ca 2+ stores, respectively. Confocal laser images were recorded from<br />

dendritic knobs of mouse ORNs loaded with the Ca 2+ -indicator dye<br />

fluo-4 AM. In OMP-nulls in response to IBMX, the rate of Ca 2+ entry<br />

to reach peak was 5-fold slower, and to achieve half-recovery to basal<br />

level was 4-fold slower compared to controls. Following KCl<br />

depolarization, the OMP-null mice showed a 2-fold delay in the halfrecovery<br />

time of the Ca 2+ peak. Similar results were obtained following<br />

Ca 2+ store depletion by stimulating ryanodine receptors with caffeine.<br />

These results suggest that the cytoplasmic Ca 2+ clearance of ORNs is<br />

significantly impaired in the OMP-null mice. These data imply that<br />

OMP is a modulator of Ca 2+ extrusion or sequestration processes in<br />

ORNs. Characterization of the mechanism by which OMP modulates<br />

Ca 2+ flux in ORNs is under investigation.<br />

Supported by NIH DC03112 (FLM), NIH DC DC00347 (FLM, FZ),<br />

NIH DC003773 (TL-Z).


151 Poster [ ] Odor Activation and Modulation<br />

OMP: A CAUTIONARY TALE OF A GENE WITHIN A GENE<br />

Margolis J.W. 1, Munger S.D. 1, Zhao H. 2, Margolis F.L. 1 1Dept Anat &<br />

Neurobiology, University of Maryland School of Medicine, Baltimore,<br />

MD; 2Dept Biology, Johns Hopkins University, Baltimore, MD<br />

Expression of the olfactory marker protein gene (Omp) is very highly<br />

restricted to mature olfactory receptor neurons and vomeronasal<br />

receptor neurons in vertebrates. A related cDNA has been cloned from<br />

the snail Eobania vermiculata (Mazzatenta et al., 2002), suggesting that<br />

OMP is of broad phylogenetic distribution, distant evolutionary origin<br />

and functional significance. The detailed characterization of the OMP<br />

gene and protein has made it a desirable locus <strong>for</strong> genetic manipulation<br />

in the olfactory system. The Omp locus, including the intronless coding<br />

region <strong>for</strong> OMP as well as all its regulatory elements, spans about 12kb<br />

in the rodent genome. To our surprise, our recent in silico analysis of<br />

this genetic region has revealed that the mouse Omp locus is entirely<br />

contained within an ~ 16kb intron of the gene encoding the Calpain 5<br />

(Capn5) protease. The Omp and and Capn5 genes are similarly<br />

organized in the human, rat and Fugu genomes. These analyses beg a<br />

critical question: do the observed phenotypic alterations that result from<br />

genetic manipulations of the Omp locus reflect (1) a direct and accurate<br />

result of these manipulations or (2) an alteration in the splicing and/or<br />

expression pattern of Capn5? We will present data that address these<br />

concerns.<br />

Supported by NIH grants DC003112(FLM), DC005633(SDM),<br />

DC006178(HZ)<br />

152 Poster [ ] Odor Activation and Modulation<br />

RESPONSE PROFILES AND NARROWING SELECTIVITY OF<br />

OLFACTORY RECEPTOR NEURONS OF XENOPUS LAEVIS<br />

TADPOLES<br />

Manzini I. 1, Schild D. 1 1Department of Molecular Neurophysiology,<br />

University of Goettingen, Goettingen, Germany<br />

In olfactory receptor neurons (ORNs) of aquatic animals amino acids<br />

have been shown to be potent stimuli. Here we report on calcium<br />

imaging experiments in slices of the olfactory mucosa of Xenopus<br />

laevis tadpoles. We were able to determine the response profiles of 283<br />

ORNs to 19 amino acids, where one profile comprises the responses of<br />

one ORN to 19 amino acids. 204 out of the 283 response profiles<br />

differed from each other. 36 response spectra occurred more than once,<br />

i.e. there were 36 classes of ORNs identically responding to the 19<br />

amino acids. The number of ORNs that <strong>for</strong>med a class ranged from 2 to<br />

13. Shape and duration of amino acid-elicited [Ca2+ ] transients showed<br />

i<br />

a high degree of similarity upon repeated stimulation with the same<br />

amino acid. Different amino acids, however, in some cases led to<br />

clearly distinguishable calcium responses in individual ORNs.<br />

Futhermore, ORNs clearly appeared to gain selectivity over time, i.e.<br />

ORNs of later developmental stages responded to less amino acids than<br />

ORNs of earlier stages. [Supported by DFG:SFB 406 (B5) and by<br />

DFG:SPP Molecular Sensory Physiology]<br />

39<br />

153 Poster [ ] Odor Activation and Modulation<br />

RESPONSES OF OLFACTORY RECEPTOR NEURONS<br />

LACKING SPONTANEOUS ACTIVITY TO AMINO ACID<br />

STIMULI IN BLACK BULLHEAD CATFISH (AMEIURUS<br />

MELAS)<br />

Valentincic T. 1, Dolensek J. 1 1Department of Biology, University of<br />

Ljubljana, Ljubljana, Slovenia<br />

The cilia and microvilli of olfactory receptor neurons (ORNs) of<br />

freshwater fish are directly exposed to water containing few ions. We<br />

showed that non-spontaneously active ORNs respond to amino acid<br />

stimuli even when exposed <strong>for</strong> several hours to highly purified water<br />

(R>18.2 megaΩcm). Using extracellular platinum black electrodes, we<br />

recorded concentration dependent responses to 10 amino acid stimuli:<br />

L-nVal, L-Met, L-Ala, L-Leu, L-Ser, L-Lys, L-Val, L-Arg, L-Ile and L-<br />

Pro at 10-4 M. Five non-spontaneously active ORNs responded to LnVal<br />

(lowest threshold, 10-7 M) and L-Met only, and two ORNs<br />

responded to L-nVal and L-Val only. The only completely specialist<br />

cell responded to L-Ala. Additional extracellular recording sites were<br />

selected by moving the electrode and searching <strong>for</strong> a response to LnVal.<br />

At 15 such locations, responses to other amino acids were<br />

recorded. Responses to L-Met were observed at 13, to L-Ala at 9, to L-<br />

Ser at 9, to L-Leu at 7, to L-Val at 4, to L-Ile at 4, to L-Arg at 4, to L-<br />

Lys at 3 and to L-Pro at 2 locations. All recording locations were<br />

equivalent and situated along lamellae 2-7. The number of locations<br />

where responses to specific amino acid stimuli were observed correlates<br />

highly with the magnitude of the EOG responses (Spearman<br />

correlation, R=0.91; P


155 Poster [ ] Odor Activation and Modulation<br />

EFFECTS OF GNRH ON TIGER SALAMANDER OLFACTORY<br />

RECEPTOR NEURON RESPONSES TO AMINO ACIDS<br />

Yurchenko O. 1, Delay R. 2, Wirsig-Wiechmann C.R. 1 1Cell Biology,<br />

University of Oklahoma Health <strong>Sciences</strong> Center, Oklahoma City, OK;<br />

2Biology, University of Vermont, Burlington, VT<br />

To assess the role of the terminal nerve in olfaction we tested the<br />

hypothesis that gonadotropin-releasing hormone (GnRH) alters<br />

olfactory receptor neuron (ORN) responses to odorants. Recordings<br />

were made from acutely isolated ORNs of tiger salamanders using the<br />

loose patch-clamp technique and yielded data <strong>for</strong> 60 ORNs. Lglutamate<br />

and L-alanine singly and as a mixture (AA) were used as<br />

stimuli. Application of the AA mixture (both amino acids) evoked<br />

hyperpolarizing responses in 41% of ORNs and depolarizing responses<br />

in 12% of ORNs; 47% of ORNs did not respond the AA mixture. A<br />

hyperpolarizing response was dominant when L-alanine alone was<br />

used. L-glutamate evoked only hyperpolarizing responses. GnRH<br />

modulated chemosensory responses of ORNs. GnRH increased the<br />

amplitude of hyperpolarizing responses to the AA mixture and single<br />

amino acids. The effect of GnRH on depolarization could not be<br />

assessed due to few cells responding with depolarization. Of the 60<br />

neurons tested, 28 did not show a response to the AA mixture or<br />

individual amino acids. In these ORNs application of GnRH with AAs<br />

induced the appearance of excitatory or inhibitory responses.<br />

Application of GnRH alone never induced responses in ORNs. These<br />

results suggest that GnRH can modulate odorant sensitivity of ORNs<br />

and may play a role in governing the balance of in<strong>for</strong>mation flow to the<br />

olfactory bulb. Supported by an Oklahoma Center <strong>for</strong> the Advancement<br />

of Science & Technology (OCAST) grant HR00-078 (CRW).<br />

156 Poster [ ] Odor Activation and Modulation<br />

GONADOTROPIN-RELEASING HORMONE (GNRH)<br />

MODULATES K+ CURRENTS IN TIGER SALAMANDER<br />

OLFACTORY RECEPTOR NEURONS<br />

Park D. 1, Eisthen H.L. 1 1Zoology, Michigan State University, East<br />

Lansing, MI<br />

We have previously shown that GnRH, a peptide present in the terminal<br />

nerve, modulates the voltage-dependent Na + current and general<br />

odorant responses in salamanders. In this study, we used whole-cell<br />

patch clamp recordings to investigate the effects of GnRH on K +<br />

currents in isolated olfactory receptor neurons from tiger salamanders<br />

(Ambystoma tigrinum). First, we demonstrated that bath application of<br />

10 µM GnRH onto olfactory receptor neurons suppresses outward<br />

currents in 42% of cells tested. We are now determining which outward<br />

currents are affected by GnRH. Substituting Ba2+ <strong>for</strong> Ca2+ to block<br />

activation of Ca2+ -dependent currents, we found that application of<br />

GnRH decreases the magnitude of Ca2+ -independent outward currents<br />

by about 50% in 60% of the cells examined. We then blocked the<br />

transient K + current ('A current') by adding 5 mM 4-aminopyrine (4-<br />

AP) to the bath in addition to Ba2+ , and found that GnRH suppresses the<br />

magnitude of the delayed rectifier K + current in 39% of the cells tested.<br />

We are now investigating the effects, if any, of GnRH on the transient<br />

K + current and on the Ca2+ -dependent K + current. In addition to the<br />

suppressive effects described here, we have found that about 10% of the<br />

cells tested display an increase in the overall outward current during<br />

GnRH application. These results demonstrate that GnRH modulates<br />

voltage-activated K + currents. Overall, terminal nerve modulation may<br />

play an important role in peripheral olfactory signal transduction. This<br />

study was conducted in accordance with PHS guidelines, and supported<br />

by NSF (IBN 9982934) and NIH (DC05366).<br />

40<br />

157 Poster [ ] Odor Activation and Modulation<br />

PUTATIVE REPRODUCTIVE PHEROMONES IN THE ROUND<br />

GOBY, NEOGOBIUS MELANOSTOMUS: BIOSYNTHESIS<br />

AND OLFACTORY MUCOSAL RESPONSES.<br />

Arbuckle W. 1, Belanger A. 2, Scott A. 3, Li W. 4, Corkum L. 1, Yun S.S. 4,<br />

Yu K. 1, Zielinski B. 1 1Biological <strong>Sciences</strong>, University of Windsor,<br />

Windsor, Ontario, Canada; 2University of Windsor, Windsor, Ontario,<br />

Canada; 3The Center <strong>for</strong> Environment, Fisheries and Aquaculture<br />

<strong>Sciences</strong>, Weymouth, Dorset, United Kingdom; 4Fisheries and Wildlife,<br />

Michigan State University, East Lansing, MI<br />

Previous studies indicate that, in the round goby Neogobius<br />

melanostomus, the reproductively mature male releases a pheromone<br />

that attracts ripe females. These studies furthermore suggest that the<br />

pheromone may be a steroid (more specifically a 5 β reduced androgen)<br />

produced by specialized glandular tissue in the testes. In the present<br />

study, in vitro incubation of the testes converted [3H]-androstenedione<br />

into (in order of abundance): 5 β -androstan-3;α -ol-11,17-dione<br />

(11keto-etiocholanolone, 11K-ETIO); 11KETIO sulfate (11K-ETIO-s);<br />

11-ketotestosterone; 5 β -androstan-3 α-ol-17-one (etiocholanolone,<br />

ETIO); 11-;β hydroxy-androstenedione; ETIO sulfate and testosterone.<br />

11K-ETIO and 11K-ETIO-s appear to be novel compounds in teleost<br />

gonads. While olfactory potency of 11K-ETIO-s has not been tested,<br />

the response threshold concentration was 0.1 nanomolar <strong>for</strong> the 11K-<br />

ETIO and ETIO sulfate, when olfactory mucosal activity was measured<br />

by electro-olfactogram (EOG). The other compounds elicited olfactory<br />

mucosal responses at concentrations greater than 0.1 nanomolar. For<br />

11K-ETIO, ETIO sulfate, and 11-ketotestosterone; EOG response<br />

magnitudes in ripe females were significantly greater than in nonovulated<br />

females. In addition, the fact that the carbon A ring of 11K-<br />

ETIO and 11K-ETIO sulfate have a 5 β -configuration (already linked<br />

with olfactory sensitivity and behavior induction in two species of<br />

gobies) makes these likely candidates pheromones in the round goby.<br />

Supported by the Michigan Great Lakes Protection Fund, NSERC<br />

and the University of Windsor Faculty of Graduate Studies


158 Poster [ ] Odor Activation and Modulation<br />

THE RABBIT MAMMARY PHEROMONE ELICITS<br />

RESPONSES IN THE MAIN OLFACTORY EPITHELIUM OF<br />

NEWBORN RABBITS AND RATS.<br />

Jakob I. 1, Litaudon P. 2, Schaal B. 1, Sicard G. 2 1Centre des <strong>Sciences</strong> du<br />

Gout, CNRS, Dijon, France; 2Neurosciences et Systemes Sensoriel,<br />

Universite Lyon I, Lyon, France<br />

As previously described newborn rabbits rely <strong>for</strong> survival on a<br />

pheromone released from the doe´s nipples (Hudson & Distel, 1983).<br />

The pheromone inducing suckling behavior in rabbit pups has been<br />

identified in rabbit milk as a single compound, namely trans-2-methyl-<br />

2-butenal (2MB2), and acted behaviorally species specific failing to<br />

induce suckling behavior in rat pups (Schaal et al, 2003). Using an air<br />

dilution olfactometer we recorded electro-olfactograms (EOG) on 18<br />

different sites of the mucosa of the nasal septum and turbinates of<br />

newborn rabbits and rats from postnatal days 2 to 11. We tested 2MB2,<br />

3 structurally related odorants, 1 with similar odor quality , and 3<br />

unrelated general odorants. Most recording sites of the septal and<br />

turbinate mucosa of rabbit and rat were responsive to all 8 odorants. All<br />

stimuli elicited typical EOGs with a rapid rising phase and a slower<br />

decline. Amplitudes to a given stimulus at a given recording site varied<br />

from animal to animal and were not age dependent. Topographic<br />

pattern <strong>for</strong> maximum sensitivity of 2MB2 were different in all<br />

individuals. 2MB2 was the most potent stimulus only in 30% of the<br />

instances tested. All recording sites sensitive to 2MB2 responded also<br />

to several other test odorants. This indicates that the olfactory mucosa<br />

of newborn rabbit and rat exhibited neither regions of specific<br />

selectivity nor a general increase in sensitivity to 2MB2. Supported by a<br />

grant from Ministere de la Recherche, ACI “Neurosciences Integratives<br />

et Computationnelles”<br />

159 Poster [ ] Environment and Human Olfaction<br />

CHILDREN´S PREFERENCES FOR TOBACCO ODOR:<br />

EFFECTS OF MATERNAL SMOKING AND MOOD STATES<br />

Forestell C.A. 1, Maggi L. 1, Mennella J.A. 1 1Monell Chemical Senses<br />

Center, Philadelphia, PA<br />

Ongoing research in our laboratory focuses on children´s hedonic<br />

judgments of and preferences <strong>for</strong> odors as a function of the context of<br />

early experience. Age-appropriate, game-like tasks that were fun <strong>for</strong><br />

children and minimized impact of language development were used to<br />

examine responses of 3- to 9-year-old children (N=296) and their<br />

mothers to a variety of odors ranging in hedonic valence and<br />

familiarity, one of which was tobacco smoke. Parental smoking, as well<br />

as the mood state of the parent, as determined by standard tests (e.g.,<br />

Beck Depression Inventory), influenced children´s preference <strong>for</strong><br />

tobacco odors. The data revealed that women who smoked liked the<br />

tobacco odor significantly more than non-smoking women. Likewise,<br />

children whose parents smoked were significantly more likely to prefer<br />

the odor of cigarette relative to neutral odors when compared to<br />

children of non-smokers. This preference <strong>for</strong> the tobacco odor was<br />

significantly less pronounced in children of smoking mothers who were<br />

depressed when compared to children of non-depressed smoking<br />

mothers. Depressed smoking mothers also scored significantly higher<br />

on the POMS Tension, Fatigue, Anger and Confusion and lower on the<br />

vigor scales when compared to non-depressed smoking mothers. These<br />

findings suggest that children's preferences <strong>for</strong> the odor of tobacco<br />

smoke are related to the emotional context in which their mothers<br />

experience tobacco. This research was supported by Pennsylvania<br />

Research Formula Fund and NIH Grants AA09523.<br />

41<br />

160 Poster [ ] Environment and Human Olfaction<br />

ANDROSTADIENONE EXPOSURE MODULATES MOOD<br />

RATINGS BUT NOT BEHAVIOR IN WOMEN<br />

Lundstrom J.N. 1, Olsson M.J. 1 1Psychology, Uppsala University,<br />

Uppsala, Sweden<br />

Exposure to the endogenous steroid androstadienone has repeatedly<br />

been shown to modulate women´s feeling of being focused in a positive<br />

direction (Lundström et al., 2003). The aim of the current study was to<br />

investigate whether exposure to a subthreshold concentration of<br />

androstadienone would also facilitate behavioral per<strong>for</strong>mance in an<br />

attention task. Thirty-seven women participated in a double-blind<br />

within-group experiment where they per<strong>for</strong>med a 20-minute tracking<br />

task measuring sustained attention. Either a male or a female ran the<br />

experiment. The experimental substance contained 250µM<br />

androstadienone. Both the experimental and control substances were<br />

masked by eugenol in order not to be perceptually discriminable.<br />

Effects on mood variables as well as attraction ratings of male faces<br />

were measured. Exposure to androstadienone modulated participant´s<br />

mood in that they felt more focused, social, happy, and less irritated.<br />

Interactions between test substance and sex of experimenter indicated<br />

that the presence of a male experimenter enhanced some of the effects<br />

in a positive direction. However, no effects on attention per<strong>for</strong>mance or<br />

attraction ratings were found.<br />

Support: The Swedish Research Council (HSFR: F0868)<br />

161 Poster [ ] Environment and Human Olfaction<br />

SMELLING A PARTNER'S CLOTHING DURING PERIODS OF<br />

SEPARATION: PREVALENCE AND FUNCTION<br />

Mcburney D.H. 1, Shoup M.L. 1, Streeter S.A. 1 1Psychology, University<br />

of Pittsburgh, Pittsburgh, PA<br />

Scattered reports in both the scientific and popular literature, as well<br />

as personal anecdotes suggest that it is not uncommon <strong>for</strong> persons to<br />

deliberately smell their sexual partner´s clothing, especially when<br />

separated. To our knowledge there has been no documentation of the<br />

frequency of this behavior. We asked undergraduate men and women<br />

who were, or had ever been, in a committed heterosexual relationship if<br />

they had ever deliberately smelled their partner´s clothing, or had slept<br />

with an article of their partner´s clothing, during periods of separation.<br />

A large majority of women had done one of these behaviors at least<br />

once; men reported doing it much less. We discuss possible functions<br />

of this behavior. It is possible individuals may be evaluating their<br />

partner´s MHC or other signals of health provided by odor, but that<br />

would not explain why the behavior occurs in the partner´s absence.<br />

Attractiveness of a partner´s odor may derive from its value in signaling<br />

the fitness benefits provided by a mate´s presence. The sex difference<br />

may reflect the fact that women especially benefit from the protection<br />

provided by a mate, or the greater importance women place on their<br />

partner´s odor (Herz & Cahill, 1997), or the greater choosiness of<br />

women in mate selection (Buss & Schmitt, 1993).


162 Poster [ ] Environment and Human Olfaction<br />

HUMAN OLFACTORY DETECTIONS OF SOCIAL AND NON-<br />

SOCIAL CHEMOSIGNALS<br />

Miller L. 1, Nomura M. 1, Umeh Y. 1, Villarreal R. 1, Chen D. 2<br />

1Psychology Department, Rice University, Houston, TX; 2Rice<br />

University, Houston, TX<br />

Olfaction is important <strong>for</strong> the survival of many animals and used <strong>for</strong><br />

detection of a wide range of social and nonsocial in<strong>for</strong>mation. Research<br />

in animals (Petrulis, et al., 1999) suggests that the mechanisms involved<br />

in processing smells serving different functions (e.g., food vs.<br />

reproduction) may be different. Previous research shows that humans<br />

can recognize individuals, distinguish between emotional states (Chen<br />

& Haviland-Jones, 2000; Chen & McClintock, in preparation), and<br />

make fine discriminations between various nonsocial smells, although<br />

few has examined olfactory sensitivities to different types of social and<br />

nonsocial smells within the same study. In this study, we investigate<br />

and compare human olfactory sensitivities to a variety of social and<br />

nonsocial chemosignals. Thirty-two college-aged adults were asked to<br />

identify themselves, their roommate, and to distinguish between<br />

different emotional states based on the smell of sweat. Their<br />

sensitivities to nonsocial smells were also assessed using established<br />

clinical tests (threshold and odor identification tests). Over three-fourths<br />

of the subjects identified themselves in a well-controlled three-item<strong>for</strong>ced-choice<br />

task (chance = 33%, p


166 Poster [ ] Environment and Human Olfaction<br />

PERITHRESHOLD NOT SUPRATHRESHOLD EXPOSURE<br />

INCREASES SENSITIVITY TO ODORS<br />

Diamond J. 1, Dalton P. 1, Breslin P.A. 1 1Monell Chemical Senses<br />

Center, Philadelphia, PA<br />

Repeated threshold test exposures to perithreshold odorants can lead<br />

to dramatic sensitivity increases among young women, (Dalton,<br />

Doolittle & Breslin, 2002), even when the target odorant is<br />

superimposed on a suprathreshold `background´ odorant (Dalton,<br />

Diamond & Breslin, 2003). To evaluate whether this phenomenon is<br />

concentration-dependent such that perithreshold stimulation is required<br />

to induce sensitization, 9 subjects (5 females, 18-24) were exposed to<br />

suprathreshold concentrations of benzaldehyde or citralva in the process<br />

of determining Weber fractions. After multiple sessions, absolute odor<br />

detection thresholds were measured <strong>for</strong> experimental and control odors.<br />

In contrast to prior studies, we found no difference in sensitivity<br />

changes between males and females: sensitivity to both control and<br />

experimental odors increased up to 1 order of magnitude <strong>for</strong> both<br />

groups. The same subjects were later exposed to perithreshold<br />

concentrations of one of the experimental odors over ten sessions.<br />

Under these conditions, females increased sensitivity by almost three<br />

orders of magnitude, whereas males increased sensitivity by only<br />

slightly more than one. This suggests that attention to suprathreshold<br />

stimulation was insufficient to elicit sensitization and that perithreshold<br />

stimuli may be required.<br />

Supported by NIH RO1 DC 03704 & RO1 DC 02995<br />

167 Poster [ ] Environment and Human Olfaction<br />

ODOR PERCEPTION AND JUDGED PROBABILITIES OF<br />

HEALTH RISK<br />

Dalton P. 1, Maute C. 1, Naqvi F. 1 1Monell Chemical Senses Center,<br />

Philadelphia, PA<br />

The purpose of the present study was to determine how intuitive<br />

notions about the significance of odor from a chemical source combine<br />

with rational, scientific in<strong>for</strong>mation to <strong>for</strong>m risk judgments in<br />

individuals. Subjects were given in<strong>for</strong>mation pertaining to an<br />

imaginary odor, including the following benchmarks: the concentration<br />

(in parts per million) at which the odor can be detected, the<br />

concentration at which the odor is considered harmless, and the<br />

concentration at which the odor is considered harmful. In addition to<br />

this in<strong>for</strong>mation, three samples of 1-butanol at varying concentrations<br />

were provided to correspond to these benchmarks of the imaginary<br />

odor. The order of the benchmarks was different <strong>for</strong> each of three<br />

groups tested. Armed with this in<strong>for</strong>mation, subjects were asked to rate<br />

the level of risk associated with exposure to several concentrations of<br />

the imaginary odor that fell above and below the benchmarks provided.<br />

Half of the subjects were given a tight range of concentrations to rate<br />

while the other half were given a wider range of concentrations.<br />

Results show that participants consistently perceived the chemical as<br />

being more harmful after it exceeded the level of odor detection than<br />

when it was below the level of detection, regardless of whether odor<br />

perception occurred above or below levels associated with toxicity.<br />

Supported by NIH RO1 DC03704-06<br />

43<br />

168 Poster [ ] Trigeminal Chemoreceptors<br />

P2X-RECEPTOR EXPRESSION AND THEIR CONTRIBUTION<br />

TO CHEMOSENSATION IN TRIGEMINAL NEURONS<br />

Spehr J. 1, Spehr M. 1, Hatt H. 1, Wetzel C. 1 1Zellphysiologie, Ruhr-<br />

Universität Bochum, Germany, Bochum, Germany<br />

The facial innervation pattern of trigeminal nerve fibres comprises<br />

the innervation of the nasal epithelium, where free trigeminal nerve<br />

endings contribute to detection and discrimination of chemical stimuli<br />

including odorants. The signal transduction mechanisms in sensory<br />

nerve endings underlying perception of chemical stimuli remain widely<br />

uncovered. Here, we characterized trigeminal ATP-activated P2Xreceptors<br />

in cultured rat trigeminal neurons and investigated their role<br />

in chemoperception. We identified a new subpopulation of neurons<br />

lacking typical nociceptive characteristics and expressing homomeric<br />

P2X2-receptors. Using a certain group of chemicals known as<br />

trigeminal stimuli we found no direct activation of trigeminal neurons,<br />

but a modulation of P2X2-receptor mediated currents. In contrast,<br />

P2X3-receptor mediated currents of nociceptive trigeminal neurons<br />

remained unaffected by the tested chemicals. There<strong>for</strong>e, we assume a<br />

functional role of the newly identified subpopulation in chemodetection<br />

of certain trigeminal stimuli.<br />

169 Poster [ ] Trigeminal Chemoreceptors<br />

SINGLE NUCLEOTIDE POLYMORPHISMS (SNPS) IN THE<br />

CAPSAICIN RECEPTOR: RELATIONSHIP TO CHEMO-<br />

SENSORY PERFORMANCE IN A PILOT SAMPLE<br />

Tarun A. 1, Shusterman D.J. 1 1Medicine, University of Cali<strong>for</strong>nia, San<br />

Francisco, Richmond, CA<br />

The capsaicin or vanilloid (VR1 / TRPV1) receptor responds to<br />

multiple noxious stimuli, including H + , heat, and capsaicin and its<br />

analogs. We recently observed marked inter-individual variation in<br />

subjective rating of capsaicin solution applied topically to the nasal<br />

mucosa in human subjects, and hypothesized that sequence variations in<br />

the TRPV1 gene could be responsible <strong>for</strong> this phenomenon. To test this<br />

hypothesis, we sampled genomic DNA from ten subjects (4 females, 6<br />

males; age range 24-56 years), all of whom had undergone trigeminal<br />

threshold testing with CO 2 (an acid producer) and six of whom had<br />

given suprathreshold irritation ratings to nasal capsaicin challenge. We<br />

sequenced five coding regions with known single nucleotide<br />

polymorphisms (SNPs) in the TRPV1 gene in each of the genomic<br />

DNA samples. Three of these SNPs resulted in a non-synonymous<br />

change in amino acid sequence of TRPV1. We found polymorphism in<br />

two of these non-synonymous SNPs. However neither of these<br />

polymorphisms correlated with either capsaicin suprathreshold rating or<br />

CO 2 threshold per<strong>for</strong>mance of the ten subjects. The only obvious<br />

chemosensory correlation was heterozygosity in one SNP resulting in a<br />

synonymous change in TRPV1 sequence in one subject with unusually<br />

low ratings of capsaicin irritancy. In this small pilot sample we<br />

demonstrated the feasibility of comparing chemosensory per<strong>for</strong>mance<br />

with sequence variation in the TRPV1 receptor gene. However, our<br />

results suggest that none of the five SNPs initially examined predicts<br />

inter-individual differences in capsaicin and CO 2 perception.


170 Poster [ ] Trigeminal Chemoreceptors<br />

CAPSAICIN SELF-SENSITIZATION IN CULTURED<br />

TRIGEMINAL NEURONS<br />

Bryant B. 1 1Monell Chemical Senses Center, Philadelphia, PA<br />

Sensitization of peripheral somatosensory neurons is one means by<br />

the amplification of the system can be modulated in certain cases of<br />

tissue damage and pathology. It is well known, <strong>for</strong> instance, that<br />

bradykinin (an endogenous algogen released on tissue injury) sensitizes<br />

somatosensory neurons to heat and capsaicin (CAP). When CAP is<br />

periodically applied to the oral mucosa at short intervals, typically 1<br />

min between stimuli, the reported irritation increased over time and has<br />

been termed sensitization (Green, 1991). Studies in which cultured<br />

trigeminal or dorsal root ganglion neurons have been exposed to<br />

repeated stimulation with CAP have only reported desensitization of<br />

subsequent responses to CAP stimulation. Because of this, it has been<br />

most tempting to attribute the observed increases in the intensity of<br />

sensory irritation to either a slow increase in the peripheral tissue<br />

concentration of CAP or to sensitization of central processes. To<br />

examine the peripheral basis of CAP self-sensitization, intracellular<br />

calcium responses of fura2-loaded rat trigeminal neurons were<br />

measured. Tested at concentrations lower (50nM - 1 uM) than<br />

previously tested (1-10 uM), CAP induced self-sensitization in a<br />

portion of neurons. At 1uM, repeated stimulation with CAP only<br />

induced the same amplitude or decremented responses. Between 50 and<br />

300 nM CAP, up to 46.2% of CAP-sensitive neurons exhibited<br />

sensitization. This suggests that a balance between sensitizing and<br />

desensitizing processes occurs in trigeminal neurons with the<br />

desensitizing processes having a higher threshold. Although narrow in<br />

concentration range, this demonstrates that peripheral self-sensitization<br />

by CAP does take place.<br />

171 Poster [ ] Trigeminal Chemoreceptors<br />

THE EFFECT OF VR1 BLOCKERS ON PERIPHERAL<br />

TRIGEMINAL NERVE RESPONSES TO IRRITANTS<br />

Allgood S. 1, Silver W.L. 1 1Biology, Wake Forest University, Winston-<br />

Salem, NC<br />

The vanilloid receptor (VR1, aka TRPV1), present on trigeminal<br />

nerve fibers, is a non-selective cation channel with high permeability<br />

<strong>for</strong> divalent cations. VR1 responds to heat, acids and capsaicin but its<br />

specific role in trigeminal stimulation by irritants has not been well<br />

studied. Our study seeks to further investigate the role of VR1 in<br />

peripheral trigeminal nerve responses. Neural recordings were obtained<br />

from the ethmoid nerves of adult rats in response to irritants delivered<br />

in solution to the nasal cavity. The ethmoid nerve responded to<br />

capsaicin, propionic acid, cyclohexanone and nicotine. When these<br />

compounds were presented with ruthenium red, an inhibitor of Ca2+ transport through membrane channels, there was a decreased response<br />

to capsaicin, propionic acid and cyclohexanone, but no change in the<br />

response to nicotine. When these compounds were presented with<br />

capsazepine, a competitive VR1 inhibitor, there was a decreased<br />

response to capsaicin and cyclohexanone, but no change in the response<br />

to nicotine and propionic acid. These data indicate that while capsaicin,<br />

cyclohexanone and propionic acid may at least partially exert their<br />

effect through VR1, nicotine does not. Additionally, these data show<br />

that while both ruthenium red and capsazepine work to block VR1<br />

receptors, they do so in different ways, as exhibited by their effect on<br />

the propionic acid response. Future work will look at additional<br />

compounds and receptor blockers to better understand the mechanisms<br />

of trigeminal nerve nociception.<br />

44<br />

172 Poster [ ] Trigeminal Chemoreceptors<br />

PERSISTENCE OF NASAL SOLITARY CHEMORECEPTOR<br />

CELLS AFTER NEONATAL CAPSAICIN TREATMENT<br />

Finger T.E. 1, Gulbransen B. 1, Böttger B. 1, Alimohammadi H. 2, Silver<br />

W.L. 2 1Rocky Mountain Taste & Smell Ctr., University of Colorado<br />

Health <strong>Sciences</strong> Center, Denver, CO; 2Biology, Wake Forest<br />

University, Winston-Salem, NC<br />

Our recent studies (Finger et al. PNAS 2003) show that nasal<br />

trigeminal chemosensitivity in mice and rats is mediated in part by<br />

solitary chemosensory cells (SCCs) distributed throughout much of the<br />

nasal respiratory epithelium. These SCCs express gustducin and T2R<br />

(bitter) taste receptors, and synapse onto peptidergic (substance<br />

P/CGRP) fibers of the trigeminal nerve that densely innervate the<br />

receptor cells. Administration of capsaicin to neonatal rat pups (0.1ml<br />

of a 1% capsaicin solution, 50 mg/kg) destroys the peptidergic<br />

trigeminal ganglion cells and effectively eliminates the trigeminal<br />

neural response to most irritants. These experiments were designed to<br />

test whether capsaicin-mediated elimination of trigeminal peptidergic<br />

innervation reduced or eliminated the nasal SCCs, i.e. whether nasal<br />

SCCs are dependent on innervation. Two-4 months following neonatal<br />

injection of capsaicin, rats were anesthetized and prepared <strong>for</strong><br />

electrophysiology. In 3 of 3 desensitized animals, the ethmoid nerve<br />

showed no responses to cycloheximide solution applied to the nasal<br />

mucosa – a solution that evokes robust activity in normal animals.<br />

Following recordings, these and other desensitized rats were perfused<br />

with 4% para<strong>for</strong>maldehyde and prepared <strong>for</strong> immunocytochemistry.<br />

Despite nearly total elimination of CGRP-immunoreactive nerve fibers,<br />

the gustducin-expressing nasal SCCs remained. Sparse non-peptidergic<br />

innervation of the epithelium remained (as assessed with PGP antisera)<br />

but these remaining nerve fibers did not <strong>for</strong>m close, embracing contacts<br />

with the SCCs.<br />

Supported by NIDCD Grant DC0060770.<br />

173 Poster [ ] Trigeminal Chemoreceptors<br />

PATTERNS OF VARIATION IN THE BEHAVIORAL<br />

RESPONSES OF RATS TO IRRITANTS AFTER NEONATAL<br />

CAPSAICIN TREATMENT<br />

Alimohammadi H. 1, Silver W.L. 1 1Biology, Wake Forest University,<br />

Winston-Salem, NC<br />

Administration of capsaicin to neonatal rat pups has been shown to<br />

produce adults with decreased trigeminal sensitivity to many irritants.<br />

This observed loss in chemosensitivity presumably occurs through the<br />

elimination of the vanilloid receptor-expressive cell population of the<br />

trigeminal ganglion, and it was hypothesized that the degree of<br />

desensitization in individual adults would vary depending on the<br />

efficacy of the initial capsaicin treatment. To test this hypothesis, a<br />

behavioral screen was developed to examine inter-animal variability in<br />

aversion responses to irritants. Adult rats injected with either capsaicin<br />

or a control solution as neonates were presented with a series of eight<br />

irritating/non-irritating chemical stimuli. Each rat´s response within the<br />

first few seconds of encounter with the stimulus source was then scored<br />

using a scale to measure aversive and/or favorable responses. Neonatal<br />

capsaicin treatment was found to result in varying degrees and differing<br />

patterns of desensitization to the five irritants tested. While most of the<br />

capsaicin-injected rats displayed diminished aversion reactions when<br />

compared to control animals, a few (3 of 13) were found to be more<br />

sensitive than the least sensitive control animal. Among the most<br />

desensitized animals, sensitivity to acetic acid remained intact whereas<br />

sensitivity to nicotine, cyclohexanone, amyl acetate, and ethanol were<br />

diminished. Capsaicin treatment had no effect on the behavioral<br />

response to non-irritating stimuli. These findings show that neonatal<br />

administration of capsaicin does not guarantee a desensitized-adult<br />

state, and may result in varying degrees of desensitization.


174 Poster [ ] Trigeminal Chemoreceptors<br />

TOPOGRAPHICAL DIFFERENCES IN THE SENSITIVITY OF<br />

THE INTRANASAL TRIGEMINAL SYSTEM<br />

Frasnelli J. 1, Heilmann S. 1, Hänel T. 1, Hummel T. 1 1University of<br />

Dresden Medical School, Dresden, Germany<br />

Preliminary data indicate that the nasal mucosa shows a differential<br />

responsiveness to trigeminal stimuli depending on the site of<br />

stimulation. Aim of the present study was the comparison of the<br />

sensitivity of two different areas of the respiratory epithelium to<br />

mechanical (air puffs) and chemosensory (gaseous CO2) stimuli.<br />

Specifically, stimuli were applied to the anterior portion of the nasal<br />

cavity and to the pharyngeal area. Responses were quantified using<br />

psychophysical (intensity) and electrophysiological (event related<br />

potentials - ERPs) techniques. For all measured parameters, a<br />

significant interaction between “location” * “stimulus” emerged,<br />

indicating that (i) CO2 evoked ERP of higher amplitude and higher<br />

intensity ratings when applied to the anterior portion of the nasal cavity<br />

compared to pharyngeal stimulation, and (ii) mechanical stimuli elicited<br />

larger responses when presented to the pharynx than to the anterior<br />

nasal cavity. These findings suggest that the nasal mucosa does not<br />

exhibit a homogenously distributed sensitivity, but that, in addition to<br />

the olfactory mucosa, there are areas with specific sensory function.<br />

These topographical differences in intranasal sensitivity may play a role<br />

in terms of differences between the ortho- and retronasal perception of<br />

odors.<br />

175 Poster [ ] Trigeminal Chemoreceptors<br />

VIRAL TRACING OF MURINE TRIGEMINAL NEURONS<br />

INNERVATING THE NASAL CAVITY<br />

Damann N. 1, Klupp B. 2, Mettenleiter T.C. 2, Hatt H. 1, Wetzel C. 1 1Cell<br />

physiology, Ruhr-University, Bochum, Germany; 2Friedrich-Loeffler-<br />

Institute, Bundes<strong>for</strong>schungsanstalt für Viruskrankheiten der Tiere, Insel<br />

Riems, Germany<br />

The trigeminal nerve is the major mediator of sensations from the<br />

mammalian head and comprises neurons that transduce mechanical,<br />

thermal and chemical stimuli. Thereby single neurons mediate sensory<br />

input from selective areas of the head (meninges, cornea and<br />

conjunctiva of the eyes, facial skin, mucous membranes of the oral and<br />

nasal cavities). Differential physiological features of peripheral neurons<br />

depending on their function and area of innervation remain largely<br />

unclear. Viral tracing was per<strong>for</strong>med to identify trigeminal neurons that<br />

mediate in<strong>for</strong>mation from the murine nasal cavity. After application of<br />

high titered Pseudorabies virus (PrV) into the nose, markerprotein and<br />

immunhistochemistry based investigations were carried out. Paraffin<br />

embedded sections and whole mount preparations were used to describe<br />

viral spread. Histochemical investigations revealed an ipsilateral spread<br />

via the ophthalmic division of the trigeminal nerve to the gasserian<br />

ganglion (GG). Infected GFP labeled ganglion neurons could also be<br />

identified after dissociation and plating allowing activity measurement<br />

of individual identified neurons in primary cell culture. PrV constitutes<br />

a powerful tool to per<strong>for</strong>m rapid tracing of the murine trigeminal<br />

system and af<strong>for</strong>ds the possibility to selectively label neurons<br />

innervating the nose. Electrophysiological and calcium imaging based<br />

characterization of these neurons concerning their specificity <strong>for</strong><br />

selected chemical compounds shall be per<strong>for</strong>med in future research.<br />

45<br />

176 Poster [ ] Trigeminal Chemoreceptors<br />

GENDER DIFFERENCES AND NASAL INTEGRATION<br />

STUDIES PERFORMED USING AN OCULAR EXPOSURE<br />

DEVICE FOR DETECTION OF IRRITATION THRESHOLDS:<br />

THE T.I.D.E. SYSTEM<br />

Opiekun R.E. 1, Mcdermott R. 1, Dalton P. 1 1Monell Chemical Senses<br />

Center, Philadelphia, PA<br />

Ocular irritation from low-level volatile chemicals is one of the most<br />

common complaints in occupational and residential environments. In<br />

contrast with techniques available <strong>for</strong> nasal irritation assessment,<br />

however, there are few standardized methods available <strong>for</strong> measuring<br />

ocular irritation thresholds. Based on the lateralization technique<br />

originally developed <strong>for</strong> nasal irritation thresholds, we have developed<br />

an instrument that can deliver controlled concentrations of a chemical<br />

vapor to one eye and a blank stimulus to the other without producing<br />

mechanical somatosensory stimulation that might alter ocular<br />

thresholds <strong>for</strong> chemicals. Initial results suggest that this system can be<br />

very useful <strong>for</strong> objective measurement of ocular irritation, that<br />

simultaneous nasal stimulation affects ocular thresholds and that<br />

females have a lower ocular threshold upon integration than males.<br />

Results also indicate that while there is a wide variation of ocular<br />

irritation thresholds among individuals, vapor concentrations required<br />

to initiate conjunctival irritation can be substantially lower than those<br />

encountered in occupational settings.<br />

177 Poster [ ] Trigeminal Chemoreceptors<br />

LATERALIZATION OF CHEMOSENSORY STIMULI:<br />

EFFECTS OF OLFACTORY FUNCTION, AGE AND GENDER<br />

ON TRIGEMINALLY MEDIATED SENSATIONS<br />

Reden J. 1, Futschik T. 1, Frasnelli J. 2, Huettenbrink K. 2, Hummel T. 3<br />

1Department of Otorhinolaryngology, University of Dresden Medical<br />

School, Dresden, Germany; 2Otorhinolaryngology, University of<br />

Dresden Medical School, Dresden, Germany; 3University of Dresden<br />

Medical School, Dresden, Germany<br />

The present investigation aimed to compare trigeminal nasal function<br />

of anosmic and hyposmic patients to healthy controls. Further, we<br />

aimed to study effects of age and gender on trigeminally mediated<br />

sensations following intranasal chemosensory stimulation. Participants<br />

were 35 patients with olfactory disfunction (n=13: functional anosmia;<br />

n= 22: hyposmia; age 28-69 years). Their results were compared with<br />

17 normosmic subjects (age 28-82 years). Olfactory function was<br />

assessed using the “Sniffin´ Sticks” test kit (butanol odor threshold,<br />

odor discrimination, odor identification). The subjects´ ability to<br />

lateralize odors was investigated <strong>for</strong> benzaldehyde and eucalyptol.<br />

Patients with olfactory dysfunction had lower scores in the<br />

lateralization task than controls (P


178 Poster [ ] Functional Organization of the Gustatory<br />

System<br />

ION SUBSTITUTION AFFECTS THE LINGUAL SURFACE<br />

POTENTIAL (LSP) IN HUMANS<br />

Feldman G. 1, Heck G. 2, Mogyorosi A. 1 1Medicine, Virginia<br />

Commonwealth University, Richmond, VA; 2Physiology, Virginia<br />

Commonwealth University, Richmond, VA<br />

We have demonstrated in humans that Na + evokes a LSP and that<br />

amiloride inhibits a portion of the Na + -evoked LSP in some but not all<br />

subjects (J. Neurophysiol. 90: 2060, 2003). To further characterize<br />

electrophysiologic aspects of the human tongue we examined whether<br />

ion substitutes influence the LSP. Initial studies employed K + and Li + ,<br />

because Li + transits Na + -selective pathways, e.g. ENaC, while K + does<br />

not readily transit such pathways. The LSP response to K + and Li +<br />

depended on whether the individual´s Na + -evoked LSP was inhibited by<br />

amiloride. In subjects who manifested amiloride sensitivity, K +<br />

substitution reduced the LSP, while Li + ´s effect was similar to that of<br />

Na + . In these subjects, after addition of amiloride to solutions, the Na + -<br />

and Li + -evoked LSPs were reduced and were similar to that of K + . In<br />

subjects who had no amiloride sensitivity, the effect of K + on the LSP<br />

was similar to those of Na + and Li + . In other studies gluconate and<br />

sulfate replaced Cl – . While the substitute anions increased the Na + -<br />

evoked LSP, the sulfate effect was approximately 3 fold greater than<br />

that of gluconate. The effects of the anion substitutes were not<br />

influenced by amiloride or the subject´s sensitivity to amiloride. These<br />

data complement our previous observation that some but not all<br />

individuals manifest amiloride sensitivity of their Na + -evoked LSP.<br />

These data also indicate that the amiloride sensitive Na + pathway<br />

allows Li + to transit, but excludes K + , consistent with the presence of<br />

ENaC in a subset of humans. Anions also influence the Na + -evoked<br />

LSP, but they do so independently of ENaC.<br />

46<br />

179 Poster [ ] Functional Organization of the Gustatory<br />

System<br />

EXPRESSION OF DELAYED RECTIFYING K CHANNELS IN<br />

TASTE CELLS OF OBESITY-PRONE AND –RESISTANT<br />

RATS.<br />

Hansen D.R. 1, Gilbertson T.A. 1 1Biology & Center <strong>for</strong> Integrated<br />

BioSystems, Utah State University, Logan, UT<br />

Direct fatty acid-mediated inhibition of delayed rectifying potassium<br />

(DRK) channels, specifically those in the Shaker subfamily (Kv1), has<br />

been proposed as a taste transduction mechanism <strong>for</strong> dietary fat. Our<br />

previous work demonstrated that fungi<strong>for</strong>m TRCs responded only to<br />

cis-polyunsaturated fatty acids (PUFAs), while those found in the<br />

foliate and vallate papillae, responded to both PUFAs and a subset of<br />

monounsaturated fatty acids. In addition we demonstrated that TRC<br />

sensitivity to PUFAs was inversely correlated with dietary fat<br />

preference. Patch clamp recording has shown that TRCs in Osborne-<br />

Mendel (OM) rats, which show a marked preference <strong>for</strong> dietary fat and<br />

become obese when placed on a high fat diet, are much less sensitive to<br />

PUFAs than are TRCs in S5B/Pl (S5B) rats that tend to avoid fat and<br />

remain lean even on a high fat diet. Our hypothesis is that TRCs in S5B<br />

rats express a greater ratio of fatty acid-sensitive to fatty acidinsensitive<br />

DRK channels than do those in OM rats. To determine if<br />

this difference in PUFA responsiveness is correlated with expression<br />

levels of DRK channels, we used real-time quantitative PCR to measure<br />

expression levels of DRK channels in taste buds from OM and S5B<br />

rats. Consistent with DRK current densities, which are twice as great in<br />

OM rats than in S5B rats, TRCs from OM rats exhibit greater<br />

expression of DRK channels. Specifically, the Kv2 (Shab) and Kv3<br />

(Shaw) DRK channel families, which may represent the PUFAinsensitive<br />

DRK channels, are much more highly expressed in OM rats<br />

than in S5B rats consistent with our hypothesis. Supported by NIH<br />

DK59611 (TAG).<br />

180 Poster [ ] Functional Organization of the Gustatory<br />

System<br />

THE PORE-FORMING ANTIBIOTIC NYSTATIN INHIBITS<br />

TASTE CELL K CURRENTS IN PERFORATED PATCH<br />

RECORDINGS.<br />

Klein J.T. 1, Guenter J. 1, Rosenthal A. 1, Gilbertson T.A. 1 1Biology &<br />

Center <strong>for</strong> Integrated BioSystems, Utah State University, Logan, UT<br />

Nystatin and amphotericin B, pore-<strong>for</strong>ming antibiotics commonly<br />

used in per<strong>for</strong>ated patch clamp recording, have been shown to inhibit<br />

and increase the rate of inactivation of the delayed rectifying K channel<br />

Kv1.3 when applied to the intracellular face of the membrane (Hahn et<br />

al. Neuropharmacology 35: 895, 1996). Since it is not known whether<br />

these antibiotics can gain access to the intracellular face and affect K<br />

currents in the per<strong>for</strong>ated patch configuration, we investigated this<br />

possibility in taste receptor cells (TRCs), which we have shown by RT-<br />

PCR and immunocytochemistry to express Kv1.3 as well as other<br />

delayed rectifying K channels. We found that nystatin, but not<br />

amphotericin, caused roughly a 40% inhibition of delayed rectifying K<br />

currents in TRCs at 20 min. and increased the rate of inactivation<br />

several fold. This difference between nystatin and amphotericin may be<br />

due to the fact that the IC 50 of amphotericin´s effects on Kv1.3 is ~18<br />

times higher than that of nystatin, while the concentrations of<br />

amphotericin we found necessary to gain good electrical access to the<br />

TRCs was only 2-3 times higher than the concentrations of nystatin.<br />

Thus, one should exert caution when using pore-<strong>for</strong>ming antibiotics in<br />

per<strong>for</strong>ated patch recordings because of possible effects on ionic<br />

currents. Supported by DC02507, DC00347, DC00353 and UAES 630<br />

(TAG).


181 Poster [ ] Functional Organization of the Gustatory<br />

System<br />

THE CELLULAR AND MOLECULAR BASIS FOR WATER<br />

TASTE IN MICE.<br />

Spray K.J. 1, Baquero A. 1, Gilbertson T.A. 1 1Biology & Center <strong>for</strong><br />

Integrated BioSystems, Utah State University, Logan, UT<br />

Recent studies have begun characterizing water (hypoosmotic<br />

stimuli) responses in the peripheral gustatory system in rat. Increases in<br />

activation were associated with rapid and reversible changes in cell<br />

volume. Such rapid changes in cell volume may be due to aquaporin<br />

channels (AQP) previously identified in taste receptor cells (TRC). We<br />

have begun examining this response in two inbred mice strains<br />

(C57/6ByJ & 129X/SvJ) using whole cell patch clamp recording.<br />

Approximately half of the fungi<strong>for</strong>m TRCs in each strain responded to<br />

a moderately hypoosmotic solution (255 mOsm) with reversible<br />

increases in conductance, presumably due to activation of a volumesensitive<br />

Cl channel. Immunocytochemistry showed apical expression<br />

of AQP5 in isolated fungi<strong>for</strong>m TRCs in both strains. We have also<br />

examined how these mice respond behaviorally to changes in solution<br />

osmolarity using mannitol (believed to be tasteless) to alter tonicity. 24h<br />

3-bottle intake tests assessed preferences <strong>for</strong> several concentrations of<br />

mannitol (55 mM-330 mM) to distilled water. As concentration<br />

(osmolarity) increased, we found decreased mannitol preference. There<br />

were also significant strain differences at lower concentrations.<br />

Preference scores <strong>for</strong> mannitol were lower <strong>for</strong> the B6 than 129 mice.<br />

There<strong>for</strong>e, mice respond to changes in osmolarity both behaviorally and<br />

at the cellular level. In addition, expression of water channels,<br />

particularly the apically expressed AQP5, may play an important role in<br />

the initial transduction events. Current studies are addressing the link<br />

between AQP expression and water responsiveness in mice. Supported<br />

by NIH DC02507 (TAG).<br />

182 Poster [ ] Functional Organization of the Gustatory<br />

System<br />

MONITORING T1R TASTE RECEPTOR DIMERIZATION<br />

Ji Q. 1, Snyder L. 2, Benard L. 1, Max M. 2, Margolskee R. 1 1Department of<br />

Physiology & Biophysics, Howard hughes medical institute, Mount<br />

Sinai School of Medicine, New York, NY; 2Department of Physiology &<br />

Biophysics, Mount Sinai School of Medicine, New York, NY<br />

T1Rs are taste cell-expressed type 3 G-protein-coupled receptors<br />

(GPCRs). Dimerization of T1R receptors has been inferred, but not<br />

actually shown. We have used BRET (Bioluminescence Resonance<br />

Energy Transfer) assay to demonstrate T1R receptor dimerization.<br />

BRET2 assay is based on energy transfer between fusion proteins<br />

containing Renilla luciferase (Rluc) and Green Fluorescent Protein<br />

(GFP).<br />

Human T1R1 (hT1R1), hT1R2 and hT1R3 coding regions were<br />

subcloned into Rluc and GFP vectors. HEK-293T cells were cotransfected<br />

with all pairs of hT1R-Rluc and hT1R-GFP plasmids.<br />

Significant BRET signals were observed in cells expressing hT1R1 &<br />

T1R3 pairs, hT1R2 & hT1R3 pairs, but not hT1R1 & hT1R2 pairs.<br />

Saturation curves were constructed <strong>for</strong> each dimer pair. BRET signals<br />

were largely independent of receptor density indicating that they came<br />

from dimerization and not random collision events. BRET competition<br />

assays with untagged hT1R3 reduced the dimer BRET signal,<br />

confirming the specificity of the interactions. We generated multiple<br />

hT1R3 mutants to investigate the effects of disulfide bonds on dimer<br />

<strong>for</strong>mation and ligand binding.<br />

47<br />

183 Poster [ ] Functional Organization of the Gustatory<br />

System<br />

EXPRESSION OF RGS IN TASTE BUD CELLS<br />

Wang H. 1, Max M. 2, Margolskee R. 3, Brand J. 1, Huang L. 1 1Monell<br />

Chemical Senses Center, Philadelphia, PA; 2Physiology and<br />

Biophysics, Mount Sinai School of Medicine, New York, NY; 3Howard<br />

Hughes Medical Institute, The Mount Sinai School Of Medicine, New<br />

York, NY<br />

Bitter, sweet and umami taste stimuli are detected by G-protein<br />

coupled receptors. The activation of these receptors in turn stimulates<br />

heterotrimeric G proteins, and triggers signal transduction cascades,<br />

eventually leading to a change in cell membrane potential and the<br />

release of neurotransmitters onto afferent axons. Recently, we and<br />

others have isolated several key taste transduction molecules. However,<br />

the mechanisms underlying the termination and desensitization of taste<br />

responses remain to be determined. To identify proteins that are<br />

involved in taste signal processing and peripheral coding, we isolated<br />

individual taste receptor cells, amplified single cell transcriptomes and<br />

constructed single taste cell cDNA libraries. By subtractively screening<br />

these libraries, we isolated a number of genes that are selectively<br />

expressed in taste cells. Among them is a regulator of G protein<br />

signaling (RGS) protein, which presumably serves as a negative<br />

regulator of heterotrimeric G-protein mediated signaling pathways by<br />

stimulating GTPase activity of the α subunits. In situ hybridization<br />

showed that this RGS is expressed in a subset of taste receptor cells.<br />

Further analysis of its interaction with taste signaling molecules in these<br />

taste cells could provide novel insights into our understanding of taste<br />

response processing and termination at the taste end organs. This work<br />

is supported by NIH grants DC05154(LH), DC03155(RFM) and<br />

MHS7241(MM), and a grant from VA(JGB).<br />

184 Poster [ ] Functional Organization of the Gustatory<br />

System<br />

DO TASTE CELLS THAT UTILIZE THE PLC SIGNALING<br />

PATHWAY ALSO EXPRESS VOLTAGE-GATED CALCIUM<br />

CHANNELS?<br />

Medler K. 1, Clapp T. 2, Sue K. 2 1Biomedical <strong>Sciences</strong>, Anatomy and<br />

Neurobiology Section, Colorado State University, Fort Collins, CO;<br />

2Biomedical <strong>Sciences</strong>, Colorado State University, Fort Collins, CO<br />

It has been well established that bitter, sweet, and umami taste<br />

stimuli are coupled to G proteins that activate the PLC signaling<br />

pathway (Zhang et al., Cell 2003). This pathway increases intracellular<br />

Ca 2+ via release from intracellular stores and store operated influx<br />

(Ogura et al., J. Neurophys. 2002). Recently, we showed that a<br />

subpopulation of these cells, the bitter-responsive, gustducin-expressing<br />

cells, lacks voltage-gated Ca 2+ channels (Medler et al., J. Neurosci.,<br />

2003). However, it is not known if lack of voltage-gated Ca 2+ channels<br />

is a characteristic of all taste cells that use this pathway. In this study<br />

we used Ca 2+ imaging to determine if any PLC-responsive cells respond<br />

to depolarization with an increase in intracellular Ca 2+ . We found that<br />

cells that respond to a PLC activator rarely show increases in<br />

intracellular Ca 2+ when depolarized with KCl (55mM). In the cells that<br />

did respond to both, the response to the PLC activator was not robust.<br />

These data suggest either some overlap between excitable cells and the<br />

PLCβ2 signaling pathway or the presence of multiple PLC iso<strong>for</strong>ms in<br />

taste cells. Further work is needed to clarify these results. Supported by<br />

DC00766 and DC 006021 to SCK.


185 Poster [ ] Functional Organization of the Gustatory<br />

System<br />

TASTE RESPONSE AND MOLECULAR EXPRESSION OF<br />

RECEPTOR CELLS OF THE MOUSE FUNGIFORM PAPILLAE<br />

Yoshida R. 1, Sanematsu K. 1, Ninomiya Y. 1 1Kyushu University,<br />

Fukuoka, Japan<br />

In taste bud cells, expression of various molecules concerned in taste<br />

reception was detected by various molecular biological techniques.<br />

Here, we investigated taste responses of receptor cells showing action<br />

potentials and mRNA expression of taste related genes in these cells at<br />

the same time. Using loose patch technique, we recorded increases in<br />

firing frequency from taste bud cells tested <strong>for</strong> taste stimuli (NaCl,<br />

Saccharin, HCl, Quinine HCl). Two types of NaCl responding cells<br />

exist: one is amiloride-sensitive and the other amiloride-insensitive. In<br />

some cells, responses to MSG were enhanced when MSG was mixed<br />

with IMP. Responses to sweet substances in some receptor cells were<br />

suppressed by apical treatment of gurmarin and recovered after apical<br />

application of β-cyclodextrin. Of 68 cells responding to taste stimuli, 40<br />

(59%) responded to one, 24 (35%) to two, and 4 (6%) to three of four<br />

taste stimuli. The entropy value presenting the breadth of<br />

responsiveness was 0.213 ± 0.252, which was close to that <strong>for</strong> the nerve<br />

fibers. These results suggest that taste cells generating action potentials<br />

have response characteristics to taste stimuli that are comparable to<br />

those <strong>for</strong> nerve fibers. After recording of taste response, single receptor<br />

cell was withdrawn from a taste bud and checked mRNA expression of<br />

taste related genes such as T1R3 by RT-PCR method. Our preliminary<br />

data indicate that a taste cell responding to sweet stimuli expressed<br />

T1R3 and gustducin mRNA. Thus, this technique might be useful to<br />

examine molecular expression in receptor cells responding to taste<br />

stimuli.<br />

186 Poster [ ] Functional Organization of the Gustatory<br />

System<br />

THE A BLOOD GROUP ANTIGEN IS EXPRESSED BY A<br />

UNIQUE SUBSET OF TASTE BUD CELLS<br />

Christy R.C. 1, Boughter J.D. 1, Smith D.V. 1 1Anatomy & Neurobiology,<br />

University of Tennessee Health Science Center, Memphis, TN<br />

Although mammalian taste cell types differ across species, most<br />

investigators recognize at least two classes of elongated, spindle-shaped<br />

cells: Type I (dark) and Type II (light) cells, based on their relative<br />

electron densities when stained with uranyl acetate. These cell types<br />

differ markedly in their morphology in transverse sections through the<br />

taste bud. A third cell type (Type III), which is electron lucent and<br />

round in transverse section like a Type II cell (and there<strong>for</strong>e a subtype<br />

of “light” cell), was first described in rabbit foliate taste buds as<br />

possessing synaptic connections with nerve fibers. Type III cells have<br />

also been described in rat and mouse vallate taste buds on the basis of<br />

specific antigen expression and ultrastructural similarity to rabbit Type<br />

III cells. Here we examine rat and mouse taste buds <strong>for</strong><br />

immunocytochemical expression of several markers that delineate<br />

differences among light cell types, which, unlike Type I cells, are<br />

heterogeneous in their expression patterns. NCAM is expressed on a<br />

subset of Type III cells and α-gustducin on a subset of Type II cells,<br />

only some of which also express the Lewisb blood group antigen. The<br />

blood group A antigen co-localizes with α-gustducin on some cells<br />

(which do not express Lewisb ) and not others but does not label any<br />

NCAM- or PGP 9.5-positive cells. Combined with the work of others<br />

showing subtypes of Type III cells expressing combinations of PGP<br />

9.5, 5-HT and/or NCAM, these data provide additional evidence <strong>for</strong> the<br />

molecular complexity of mammalian taste bud cells. Supported by<br />

NIDCD DC00347 to DVS.<br />

48<br />

187 Poster [ ] Functional Organization of the Gustatory<br />

System<br />

PROLIFERATION OF LINGUAL MACROPHAGES AFTER<br />

UNILATERAL DENERVATION OF FUNGIFORM TASTE<br />

BUDS.<br />

Mccluskey L.P. 1, Rigsby C.S. 1 1Physiology, Medical College of<br />

Georgia, Augusta, GA<br />

Within days after unilateral chorda tympani nerve (CT) section,<br />

activated ED1+ macrophages are increased on both the sectioned and<br />

intact sides of the tongue. Activated macrophages release a variety of<br />

cytokines and growth factors, which are proposed to affect sodium taste<br />

function after neural injury. We hypothesized that local proliferation of<br />

ED1+ macrophages accounts <strong>for</strong> the dramatic increase observed after<br />

nerve section. SD specified pathogen-free rats received unilateral CT<br />

section or sham section on day 0. On day 1 or 2 post-section, rats were<br />

given an injection of BrdU (50 mg / kg i.p.) 6 hr prior to sacrifice.<br />

Paraffin sections were processed <strong>for</strong> double immunofluorescent staining<br />

with antibodies to BrdU and ED1. As previously observed, there was an<br />

increase in ED1+ macrophages at both day 1 and day 2 post-sectioning.<br />

However, while BrdU+ cells were also numerous throughout tongue,<br />

few cells were double-positive. There<strong>for</strong>e, ED1+ macrophages do not<br />

appear to proliferate in the lingual environment in response to<br />

denervation. We next examined whether the increase in ED1+<br />

macrophages might be due to proliferation of resting, resident ED2+<br />

macrophages that convert to an activated phenotype. Yet there were<br />

also few ED2+ / BrdU+ cells after nerve section, suggesting that<br />

proliferation of ED2+ macrophages does not account <strong>for</strong> the increased<br />

ED1+ population at these time points. We are currently examining the<br />

possibility that circulating ED1+ cells enter the tongue in response to<br />

denervation. Resolving this issue will increase our understanding of<br />

potential immune mechanisms that regulate peripheral taste function.<br />

Supported by NIH grant 1 R01 DC005811-01A1.<br />

188 Slide [ ] Beidler Colloquium on Taste Transduction<br />

THE MAMMALIAN AMILORIDE-INSENSITIVE (AI) NON-<br />

SPECIFIC SALT TASTE RECEPTOR IS A VANILLOID<br />

RECEPTOR-1 (VR-1) VARIANT<br />

Lyall V. 1, Heck G.L. 1, Vinnikova A.K. 2, Ghosh S. 2, Phan T.T. 1, Bigbee<br />

J.W. 3, Desimone J.A. 1 1Physiology, Virginia Commonwealth<br />

University, Richmond, VA; 2Internal Medicine, Virginia Commonwealth<br />

University, Richmond, VA; 3Anatomy and Neurobiology, Virginia<br />

Commonwealth University, Richmond, VA<br />

The AI non-specific salt taste receptor is the predominant transducer<br />

of salt taste in some mammalian species, including humans. The<br />

physiological, pharmacological and biochemical properties of the AI<br />

salt taste receptor were investigated by RT-PCR, direct measurement of<br />

unilateral apical Na + fluxes in polarized rat fungi<strong>for</strong>m taste receptor<br />

cells (TRCs), and chorda tympani (CT) nerve recordings to lingual<br />

stimulation with NaCl, KCl, NH Cl and CaCl , in both the rat model<br />

4 2<br />

and the VR-1 knockout mouse model. We report that the AI salt taste<br />

receptor is a constitutively active non-selective cation channel derived<br />

from the VR-1 gene. It accounts <strong>for</strong> all of the AI CT response to Na +<br />

salts and part of the response to K + , NH4 + , and Ca2+ salts. It is activated<br />

by vanilloids (resiniferatoxin and capsaicin) and temperature (>38oC), and is inhibited by VR-1 antagonists (capsazepine and SB-366791). In<br />

the presence of vanilloids, external H + and ATP lower the temperature<br />

threshold of the channel. This allows <strong>for</strong> increased salt taste sensitivity<br />

without an increase in temperature. VR-1 knockout mice demonstrate<br />

no functional AI salt taste receptor and no salt taste sensitivity to<br />

vanilloids and temperature. We conclude that the mammalian AI nonspecific<br />

salt taste receptor is a VR-1 variant. Supported by NIDCD<br />

Grants DC-02422 and DC-00122.


189 Slide [ ] Beidler Colloquium on Taste Transduction<br />

RESIDUAL RESPONSES TO BITTER, SWEET AND UMAMI<br />

COMPOUNDS IN TRPM5 KNOCKOUT MICE<br />

Damak S. 1, Rong M. 2, Kokrashvilli Z. 2, Yasumatsu K. 3, Glendinning<br />

J.I. 4, Ninomiya Y. 3, Margolskee R.F. 2 1Nestle Research Center,<br />

Lausanne, Switzerland; 2Physiology and Biophysics, Mount Sinai<br />

School of Medicine, New York, NY; 3Kyushu University, Fukuoka,<br />

Japan; 4Barnard College, Columbia University, New York, NY<br />

To determine the role of Trpm5 in taste responses in vivo, we<br />

generated Trpm5 knockout (KO) mice by homologous recombination in<br />

ES cells. Two bottle preference tests of wildtype (WT) vs. Trpm5 KO<br />

mice showed reduced, but not abolished, responses of the KO mice to<br />

the bitter compounds quinine and denatonium. By this test the Trpm5<br />

KO mice were indifferent to concentrations of quinine and denatonium<br />

up to 1 mM and 10mM, respectively, then showed strong avoidance to<br />

higher concentrations. A brief access test using a gustometer confirmed<br />

that high concentrations of denatonium were aversive to the KO mice.<br />

The KO mice were indifferent to low concentrations of denatonium and<br />

to all concentrations of quinine. With sweet compounds, two bottle<br />

preference tests of Trpm5 KO mice showed reduced preference <strong>for</strong><br />

sucrose and the artificial sweetener SC45647. However, in the brief<br />

access test, Trpm5 KO mice were indifferent to these compounds. Postingestive<br />

effects (two bottle preference test) and/or lower sensitivity<br />

(gustometer) may affect the behavioral responses elicited by these<br />

behavioral tests and explain this apparent discrepancy. Gustatory nerve<br />

recordings are in progress to further investigate this. The Trpm5 KO<br />

mice showed residual preference <strong>for</strong> 100mM and 300mM MSG<br />

(umami) by two bottle preference test, but not by lickometer. Both<br />

behavioral tests showed that the Trpm5 KO mice avoid 1 M MSG. The<br />

behavioral responses to HCl elicited by either test were identical in the<br />

KO and WT animals. Together, these results indicate that Trpm5 is an<br />

important component of the taste response to bitter, sweet and umami<br />

compounds, but that additional Trpm5-independent response pathways<br />

exist.<br />

49<br />

190 Slide [ ] Beidler Colloquium on Taste Transduction<br />

DUAL REGULATION OF THE TASTE TRANSDUCTION ION<br />

CHANNEL TRPM5 BY CA2+ AND PIP2<br />

Liman E.R. 1, Liu D. 1 1Neurobiology, University of Southern Cali<strong>for</strong>nia,<br />

Los Angeles, CA<br />

The transduction of taste is a fundamental process that allows<br />

animals to discriminate nutritious substances from those that are<br />

noxious. Three taste modalities, bitter, sweet and amino acid, are<br />

mediated by G-protein coupled receptors. Recent evidence suggest that<br />

these receptors signal through a common transduction cascade:<br />

receptors activate phospholipase C (PLC), leading to a breakdown of<br />

phosphatidyinositol-4,5-bisphosphate (PIP ) into diacylglycerol (DAG)<br />

2<br />

and inositol 1,4,5-trisphosphate (IP ), which causes release of Ca 3 2+<br />

from intracellular stores. The ion channel, TRPM5, is essential <strong>for</strong><br />

transduction of bitter, sweet and amino acid tastes, but the mechanisms<br />

which it is activated are not well understood. We find that when<br />

expressed heterologously in HEK 293, TRPM5 <strong>for</strong>ms an outwardly<br />

rectifying cation channel that is activated downstream of Gq-coupled<br />

membrane receptors. In excised patches from CHOK1 cells expressing<br />

TRPM5, channel activity was induced by micromolar concentrations of<br />

intracellular Ca2+ , but not by DAG or IP . Sustained exposure to Ca 3 2+<br />

desensitized TRPM5 channels. Application of the membrane<br />

phospholipid, PIP enhanced the current following desensitization, but<br />

2<br />

not be<strong>for</strong>e desensitization, suggesting that loss of PIP may underlie<br />

2<br />

desensitization. These data allow us to propose a model <strong>for</strong> G proteincoupled<br />

taste transduction in which Ca2+ serves as the second<br />

messenger linking receptor signaling to membrane depolarization.<br />

Supported by grants from the NIDCD (DC04564 and DC05000).<br />

191 Slide [ ] Beidler Colloquium on Taste Transduction<br />

DEORPHANIZATION AND FUNCTIONAL SNP ANALYSIS OF<br />

TAS2R BITTER TASTE RECEPTORS<br />

Bernd B. 1, Christina K. 1, Winnig M. 1, Slack J. 2, Breslin P.A. 3, Reed<br />

D.R. 4, Tharp C.D. 4, Kim U. 5, Drayna D. 5, Meyerhof W. 1 1Molecular<br />

Genetics, German Institute of Human Nutrition Potsdam-Rehbruecke,<br />

Nuthetal, Germany; 2<strong>Givaudan</strong> Flavors Corp., Cincinnati, OH; 3Monell<br />

Chemical Senses Center, Philadelphia, PA; 4Monell Chemical Senses<br />

Center, Philadelphia, OH; 5NIDCD/NIH, Rockville, MD<br />

Taste sensitivities to different bitter compounds vary between<br />

individuals. The most prominent example <strong>for</strong> genetically determined<br />

taste differences is the bitter taste of phenylthiocarbamide (PTC) and 6-<br />

N-propylthiouracil (PROP). A recently published study (Kim et al.,<br />

Science 299, 1221-1225) showed a good correlation between the human<br />

PTC taster and non-taster phenotype and three single nucleotide<br />

polymorphisms (SNPs) in the hTAS2R38 gene, suggesting that these<br />

SNPs might cause the individual taste differences <strong>for</strong> PROP, PTC, and<br />

various other chemically related bitter compounds.<br />

Using functional expression and calcium imaging (Bufe et al., Nature<br />

Genetics 32, 397-401) we demonstrated that the hTAS2R38 taster<br />

variant can be activated by PTC. In addition we identified ligands <strong>for</strong><br />

four other human bitter taste receptors hTAS2R10, hTAS2R16,<br />

hTAS2R43, and hTAS2R44. Analysis of the human genome SNP<br />

database and own sequencing ef<strong>for</strong>ts so far revealed 14 non-conserved<br />

SNPs in the coding region of these receptors. Using site directed<br />

mutagenesis we created corresponding receptor variants and are<br />

studying their effects in our heterologous expression system. Initial<br />

results revealed that several of the hTAS2R38 SNPs altered the receptor<br />

function.


192 Symposium [ ] Olfaction and Neurodegenerative<br />

Disorders<br />

OLFACTION AND NEURODEGENERATIVE DISORDERS<br />

Doty R. 1 1Smell and Taste Center, University of Pennsylvania,<br />

Philadelphia, PA<br />

Numerous major neurological disorders are associated with olfactory<br />

dysfunction. Indeed, the first clinical sign of such diseases as<br />

Alzheimer´s disease and idiopathic Parkinson´s disease appears to be<br />

decreased olfactory function. In the case of Parkinson's disease, the<br />

prevalence of smell loss is greater than the prevalence of tremor, and is<br />

essentially equivalent to that of the other major motoric signs of the<br />

disease. Importantly, not all neurological or neurodegenerative<br />

disorders are associated with smell loss. Hence, olfactory testing can<br />

aid in the differential diagnosis of a number of disorders commonly<br />

confused on initial presentation, including Alzheimer's disease vs.<br />

major affective disorder, and Parkinson's disease vs. progressive<br />

supranuclear palsy. The olfactory evaluation of at-risk individuals,<br />

including relatives, may prove to be of considerable clinical value <strong>for</strong><br />

initiating neuroprotective therapy early in disease development. This<br />

symposium provides an up-to-date overview of relationships between<br />

olfactory function and four major neurological diseases. The<br />

participants and discussants represent active researchers in this field,<br />

and bring a variety of perspectives and a wealth of in<strong>for</strong>mation to this<br />

topic.<br />

193 Symposium [ ] Olfaction and Neurodegenerative<br />

Disorders<br />

LONGITUDAL EVALUATION OF OLFACTORY FUNCTION<br />

IN ALZHEIMER'S DISEASE<br />

Devanand D.P. 1, Tabert M.H. 1 1Psychiatry, Columbia<br />

University/NYSPI, New York, NY<br />

We evaluated the predictive utility of olfactory deficits in patients<br />

with minimal to mild cognitive impairment (MMCI). 150 outpatients<br />

presenting to a Memory Disorders Center were recruited and followed<br />

at 6-month intervals. 63 group-matched controls were followed<br />

annually. Patients had a mean age of 67 (SD 9.8) and mean education of<br />

15 (SD 4.2) years. Mean age <strong>for</strong> controls was 65 (SD 9.3) and mean<br />

education was 16 (SD 2.6). UPSIT scores were lower in patients (mean<br />

31 SD 6.4) than controls (mean 34 SD 4.2). Baseline UPSIT scores<br />

were lower in converters to Alzheimer´s Disease (AD; n=35, mean 26<br />

SD 8.2) than non-converters (mean 33, SD 4.7) (mean follow-up=39<br />

months). In Cox regression analyses, low olfaction scores predicted AD<br />

(Wald chi2=15.9, p < .0001). The UPSIT score remained a significant<br />

predictor (p < .0003, relative risk=0.9 per UPSIT point change) even<br />

when age, sex, and education were included as covariates, and when<br />

MMS was included as a covariate. ApoE e4 genotype did not differ<br />

between non-converters (22%) and converters (28%). UPSIT scores<br />

were correlated with right hippocampal volume (r=0.20, p < .03). In<br />

Cox regression analyses, low olfaction scores predicted outcome (Wald<br />

chi2=10.6, p < .002, relative risk=0.92 per point change) even when<br />

hippocampal volume, age, and gender were included. Olfactory deficits<br />

predicted conversion to AD in patients with MMCI, even after<br />

controlling <strong>for</strong> demographic, cognitive, and brain volumetric predictors.<br />

In addition, odor-induced fMRI activation is also being examined in<br />

controls and AD and MMCI patients.<br />

Grants: AG17761 (P.I. Devanand), K01AG21548 (P.I. Tabert) and<br />

from the Alzheimer´s <strong>Association</strong> (P.I. Devanand)<br />

50<br />

194 Symposium [ ] Olfaction and Neurodegenerative<br />

Disorders<br />

OLFACTORY SYSTEM DYSFUNCTION IN SCHIZOPHRENIA<br />

Moberg P.J. 1, Turetsky B.I. 1 1Psychiatry, University of Pennsylvania,<br />

Philadelphia, PA<br />

Psychophysical studies of olfaction have documented impairments in<br />

odor detection thresholds, identification, memory and hedonics in<br />

patients with schizophrenia. Similar deficits exist in non-ill family<br />

members. Despite evidence of behavioral impairment, there is little<br />

evidence that the neuroanatomical or neurophysiological substrates of<br />

olfaction are abnormal.<br />

We have conducted a series of studies examining the structural and<br />

functional integrity of the olfactory system in patients and their firstdegree<br />

relatives. 1)Using acoustic rhinometry, we have shown a<br />

reduction of the left-posterior nasal cavity in male patients. 2)Using<br />

high-resolution MRI, we found bilateral olfactory bulb volume<br />

reductions in patients, and unilateral right sided reductions in family<br />

members. 3)Following parcellation of the anterior ventromedial<br />

temporal lobe into perirhinal, entorhinal and temporal polar regions,<br />

patients exhibited selective gray matter volume reductions in the<br />

perirhinal and entorhinal areas bilaterally. 4)Using air-dilution<br />

olfactometry, both patients and non-ill family members were observed<br />

to exhibit dose-dependent abnormalities in the olfactory evoked<br />

potential.<br />

There are thus primary structural and functional abnormalities in the<br />

olfactory system, which underlie the behavioral deficits seen in<br />

schizophrenia. These abnormalities appear to denote an endophenotypic<br />

genetic vulnerability marker, rather than an index of either clinical<br />

disease status or treatment. A greater understanding of the neurobiology<br />

of these olfactory deficits could offer clues to both the basic<br />

neuropathology and the genetic precursors of this disorder.<br />

Supported by MH63381(PJM), MH59852(BIT), & NARSAD(PJM).


195 Symposium [ ] Olfaction and Neurodegenerative<br />

Disorders<br />

OLFACTORY DYSFUNCTION IN MULTIPLE SCLEROSIS<br />

Doty R.L. 1 1Smell and Taste Center, Department of<br />

Otorhinolaryngology: Head and Neck Surgery, University of<br />

Pennsylvania, Philadelphia, PA<br />

For many years it was assumed that multiple sclerosis (MS) was<br />

unaccompanied by olfactory dysfunction. While several early studies,<br />

including a pioneering study by Ansari in 1976, failed to find any<br />

olfactory deficits in this disease, more recent studies from our group<br />

have clearly demonstrated that some MS patients exhibit olfactory loss.<br />

In 1984 we reported that 23% of a group of 31 MS patients exhibited<br />

smell dysfunction, as measured by the University of Pennsylvania<br />

Smell Identification Test (UPSIT). A decade later we presented clinical<br />

cases in which smell loss was a presenting sign of MS. In a subsequent<br />

series of studies in the late 1990´s, we demonstrated that such loss was<br />

correlated with the number of MS-related plaques in subtemporal and<br />

subfrontal brain regions, but not in other brain regions. Moreover, we<br />

showed that olfactory function waxes and wanes longitudinally in<br />

accordance with the number of active plaques in these brain regions. In<br />

this presentation, I review these studies, as well as initial results of a 5year-long<br />

study of 73 MS patients and controls, in which a range of<br />

sensory and neuropsychological tests were administered, including ones<br />

tapping olfactory (i.e., odor identification, detection & discrimination),<br />

gustatory (e.g., regional taste identification), auditory (e.g., pure tone<br />

thresholds, competing words, auditory figure-ground), vestibular (e.g.,<br />

Rhomberg eyes open & closed tasks), and cognitive (e.g., Wechsler<br />

Memory Scale-Revised; National Adult Reading Test; Wisconson Card<br />

Sorting Test) functions.<br />

Supported by grant RO1 DC 02974 from the National Institute on<br />

Deafness and Other Communication Disorders.<br />

196 Symposium [ ] Olfaction and Neurodegenerative<br />

Disorders<br />

OLFACTION IN PARKINSONISM<br />

Hawkes C. 1 1Essex Neuroscience Centre, Essex, United Kingdom<br />

There has been gradual increase of interest in olfaction since it was<br />

realised that anosmia was a common feature of idiopathic Parkinson's<br />

disease (IPD) and Alzheimer-type dementia (AD). It is an intriguing<br />

possibility that the first sign of a disorder hitherto regarded as one of<br />

movement or cognition may be that of disturbed smell sense. In this<br />

review of parkinsonian syndromes the following observations can be<br />

made: 1) olfactory dysfunction is frequent and often severe in IPD and<br />

may precede motor signs of IPD 2) normal smell identification in IPD<br />

is rare and should prompt review of diagnosis unless the patient is<br />

female with tremor dominant disease. 3) olfactory impairment in<br />

suspected progressive supranuclear palsy or corticobasal degeneration<br />

is atypical and should likewise provoke diagnostic review. 4) impaired<br />

smell sense is seen in some patients at 50% risk of parkinsonism 5)<br />

biopsy of olfactory nasal neurons reveals non-specific changes in IPD<br />

and at present will not aid diagnosis.<br />

51<br />

197 Poster [ ] Development of the Gustatory System<br />

EGF SIGNALING IN PATTERNING FUNGIFORM PAPILLAE<br />

IN EMBRYONIC RAT TONGUE<br />

Liu H. 1, Mistretta C. 1 1School of Dentistry, University of Michigan, Ann<br />

Arbor, MI<br />

In previous experiments to understand how papillae <strong>for</strong>m in patterns<br />

on the embryonic tongue, we demonstrated roles <strong>for</strong> the morphogen,<br />

sonic hedgehog (Shh) in fungi<strong>for</strong>m papilla development. Fungi<strong>for</strong>m<br />

papillae <strong>for</strong>m in doubled numbers and atypical posterior locations in<br />

tongue cultures when Shh signaling is selectively disrupted with the<br />

alkaloid, cyclopamine. To further understand molecular regulation and<br />

interactions in papilla <strong>for</strong>mation, we are studying effects of epidermal<br />

growth factor, EGF. A polyclonal antibody to EGF receptor (EGFR)<br />

was used to immunolocalize EGFR in rat embryo tongues at gestational<br />

days (E) 13-18. Also, in whole tongue cultures begun at E14, we<br />

examined effects of exogenous EGF on fungi<strong>for</strong>m papilla development<br />

and on the cyclopamine-induced change in papilla patterning. EGFR<br />

localized in all layers in the dorsal epithelium of embryonic rat tongue<br />

at all ages examined. Immunoproducts were weak and homogeneous in<br />

the tongue epithelium at E13-E14. Signals became progressively more<br />

intense in the inter-papilla space, but not within the papilla epithelium,<br />

in association with papilla development from E15-E18. In E14+2 day<br />

cultures, EGF dose-dependently decreased the number of fungi<strong>for</strong>m<br />

papillae while stimulating epithelial proliferation. Moreover, preincubation<br />

with EGF, followed by culture with EGF plus cyclopamine,<br />

prevented the cyclopamine-induced change in papilla pattern. The data<br />

demonstrate that EGF signaling plays an important role in spacing of<br />

fungi<strong>for</strong>m papillae in embryonic tongue, and interacts with Shh<br />

signaling in regulating development of papilla pattern. Supported by<br />

NIH NIDCD Grant DC000456 to CMM.<br />

198 Poster [ ] Development of the Gustatory System<br />

TASTE BUD PRIMORDIA DEVELOP IN RODENT TONGUE<br />

CULTURES<br />

Walters E. 1, Mbiene J. 2 1Department of Biochemistry and Molecular<br />

Biology, Howard University, Washington, DC; 2Department of<br />

Biomedical <strong>Sciences</strong>, Texas A&M University System, Dallas, TX<br />

Lingual taste buds <strong>for</strong>m within papillae in mammals. We are<br />

studying the molecular aspects of taste bud differentiation during<br />

development. Keratin 8 is expressed in epithelial cell clusters in<br />

distinctive patterns across the developing tongue in mouse embryos<br />

be<strong>for</strong>e the innervation of tongue epithelium. To learn whether keratin 8<br />

expression proceeds in vitro as in vivo, we examined the expression<br />

patterns of keratin 8 in mouse tongue cultures. Tongue or lower jaw<br />

explants were taken from gestational days 11.5 and 12.5 mouse<br />

embryos and maintained at liquid-gas interface in conventional organ<br />

cultures. Explanted tongues were harvested after 1-6 days in culture and<br />

assayed <strong>for</strong> keratin 8 by immunocytochemistry. In E11.5 and 12.5<br />

mouse tongue explants, keratin 8 is broadly expressed in the lingual<br />

epithelium, as in vivo. Over the next 24-48 hours, keratin 8 becomes<br />

restricted to clusters of cells located in the center of epithelial placodes.<br />

Surprisingly, keratin 8 continues to be expressed in few clusters after 6<br />

days of culture. These results establish that much of taste bud<br />

development is nerve dependent in mammals. I am currently examining<br />

the gene repertoire expressed in tongue placodes that may possibly play<br />

a role in the determination of taste buds.Grant sponsor: NIDCD; Grant<br />

number: DC03503


199 Poster [ ] Development of the Gustatory System<br />

IDENTIFICATION OF DEVELOPMENTALLY REGULATED<br />

GENES EXPRESSED IN TASTE BUDS<br />

Iwatsuki K. 1, Watanabe A. 2, Aburatani H. 2, Margolskee R. 3 1Physiology<br />

& Biophysics, Mount Sinai School of Medicine, New York, NY;<br />

2RCAST, The University of Tokyo, Tokyo, Japan; 3Howard Hughes<br />

Medical Institute, Mount Sinai School of Medicine, New York, NY<br />

During their differentiation and development, taste cells express a<br />

variety of molecules in spatially and temporally regulated fashion.<br />

Although several markers selectively expressed in mature taste cells<br />

have been identified, few markers of developing taste cells are known,<br />

making it difficult to study how taste cells develop and mature from<br />

stem cells and precursor cells.<br />

To search <strong>for</strong> genes selectively expressed during mouse taste cell<br />

development we used Representational Difference Analysis (RDA) and<br />

Affymetrix gene chip analysis to compare mRNAs from circumvallate<br />

(CV) papillae ("taste") vs. surrounding non-sensory lingual epithelial<br />

tissue ("non-taste"). Most of the genes cloned by RDA, including small<br />

proline rich proteins and mesothelin, were also found by Affymetrix<br />

gene chips. Selective expression of these genes was confirmed by<br />

semi-quantitative RT-PCR. Those genes confirmed to be differentially<br />

expressed in taste vs. non-taste were examined by in situ hybridization.<br />

To date, we have identified more than 20 genes, including Trpm5 and G<br />

gamma 13, that are selectively expressed in adult taste cells. We<br />

carried out in situ hybridization on a series of CV sections from early<br />

through mature stages of taste bud development to examine temporal<br />

and spatial patterns of expression of these genes. We have identified<br />

several marker genes <strong>for</strong> taste cells which were expressed in one of two<br />

patterns: 1) early through late stages, 2) only in late stage of taste cell<br />

development.<br />

200 Poster [ ] Development of the Gustatory System<br />

RELATIONSHIP BETWEEN EXPRESSION OF<br />

POSTSYNAPTIC DENSITY PROTEIN 95 (PSD-95) AND THE<br />

DEVELOPMENT OF TASTE BUDS IN THE<br />

CIRCUMVALLATE PAPILLAE OF RAT COMPARED WITH<br />

G-GUSTDUCIN AND PROTEIN GENE PRODUCT 9.5.<br />

Ueda K. 1, El Sharaby A. 2, Wakisaka S. 2 1Psychology, University of<br />

Virginia, Charlottesville, VA; 2Dept. of Oral Anat.& Deverop. Biol.<br />

Grad. School of Dentistry, Osaka University, Suita, Osaka, Japan<br />

It is well known that neurotrophic factors are necessary <strong>for</strong><br />

<strong>for</strong>mation, development and maintenance of mammalian taste buds, and<br />

that type III taste cells have synaptic connections with gustatory nerves.<br />

However it is unclear how synapses develop in taste buds. In this study<br />

we examined the distribution of postsynaptic density protein 95 (PSD-<br />

95) immunohystochemically in circumvallate papillae (CVP) in rats<br />

aged 0, 1, 2, 4, 5, 7, 14, 21, 28 days postnatal and in adults. At<br />

adulthood, strong PSD-95-like immunoreactivity (PSD-95-LI) existed<br />

around the taste pore and some of them overlapped with neuron and<br />

type III taste receptor cell specific gene, protein gene product 9.5 (PGP<br />

9.5). Type II cells are typically viewed as making contacts with nerve<br />

fibers, but they do not make classical synapses. However we found that<br />

some cells having Gα-gustducin-LI, considered as a marker of type II<br />

cells, also had PSD-95-LI. Developmentally, many taste buds existed in<br />

the CVP at birth, but PSD-95-LI did not occur until postnatal day 5.<br />

Thereafter, the density of PSD-95-LI increased progressively until it<br />

was similar to adults between PN 21-28 days. These results suggest that<br />

there may be subtypes of receptor cells based on the presence of<br />

synapses. Further, functional synaptic contacts in the CVP do not<br />

mature until after 3rd postnatal week.<br />

52<br />

201 Poster [ ] Development of the Gustatory System<br />

NEUROTROPHIC FACTORS REGULATE THE SENSITIVITY<br />

OF GENICULATE AXONS TO SEMA3A DURING<br />

INNERVATION<br />

Saldanha J. 1, Vilbig R. 1, Rochlin M.W. 1 1Biology, Loyola University of<br />

Chicago, Chicago, IL<br />

Geniculate axons begin to penetrate the rat dorsal lingual epithelium<br />

at E17, followed by trigeminal axons (Farbman & Mbiene, J Comp<br />

Neurol 306:172-186), despite the presence of epithelial Sema3A<br />

mRNA. Trigeminal axons are repelled by Sema3A in vitro through E18<br />

(Dillon et al, ibid, 470:13-24), suggesting that Sema3A repellent<br />

activity or sensitivity to Sema3A is regulated in vivo. The neurotrophic<br />

factors that stimulate geniculate axon outgrowth in collagen gels and<br />

the response of this outgrowth to Sema3A during late embryonic stages<br />

are not known. BDNF, NT4, GDNF, and NGF (15-25 ng/ml) stimulate<br />

axon outgrowth above control media levels, although the NGF effect<br />

was small. BDNF stimulated the most rapidly advancing outgrowth,<br />

NT4 the slowest. In contrast to our results with the E18 trigeminal<br />

ganglion, Sema3A-transfected COS7 cells did not repel BDNF- (N=14)<br />

and GDNF- (N=9) stimulated outgrowth compared to control<br />

transfected COS7 cells (N=10 <strong>for</strong> each). However, NT4-stimulated<br />

outgrowth (N=17) was repelled (p


203 Slide [ ] Development of the Gustatory System<br />

TASTE PLACODES ARE PRIMARY TARGETS OF EARLY<br />

GENICULATE BUT NOT TRIGEMINAL PERIPHERAL<br />

NERVE ENDINGS IN THE DEVELOPING TONGUE OF<br />

MOUSE EMBRYOS.<br />

Mbiene J. 1 1Department of Biomedical <strong>Sciences</strong>, Texas A&M<br />

University System, Dallas, TX<br />

The establishment of a reproducible pattern of connectivity is as<br />

essential to neuron´s function as its physiological properties. Until<br />

recently innervation of gustatory papillae was thought to occur after<br />

their initial morphogenesis. However, new results from mammals<br />

indicate that much of molecular markers of papilla development are<br />

expressed prior to their overt <strong>for</strong>mation. Given that markers of taste bud<br />

genesis begin to be expressed be<strong>for</strong>e papilla morphogenesis or its<br />

completion, we hypothesized that lingual epithelial placodes (which<br />

later will harbor the embryonic taste buds) are specific targets <strong>for</strong> taste<br />

afferents. To test this idea, we traced chorda tympani and lingual nerves<br />

with different lipophilic dyes and compared the patterns of their<br />

terminal arbors in tongues of mouse embryos from E11.5 to 14.5. The<br />

results indicate that chorda tympani, in contrast to lingual nerve fibers,<br />

pioneer the tongue at E11.5. Over the next 24 hours, at E12.5, the<br />

tongue epithelium is still not contacted by nerves but the ramification<br />

patterns of both nerves in the tongue are identical. At E13.5 (when the<br />

first taste buds are apparent) chorda tympani axons make contacts with<br />

the epithelial placodes under which they exhibited distinctive terminal<br />

arbors. At E14.5 (when the fungi<strong>for</strong>m papillae first <strong>for</strong>m) chorda<br />

tympani enter and reach the apical epithelium of the papilla where the<br />

embryonic taste buds <strong>for</strong>m. These results indicate that taste buds <strong>for</strong>m<br />

at the same time as they are innervated and will be discussed in the<br />

context of the current view of non-neural induction of taste buds. Grant<br />

sponsor: NIDCD; Grant number: DC03503<br />

204 Poster [ ] Development of the Gustatory System<br />

ROLES FOR HEDGEHOG PROTEINS IN SUPPORTING<br />

NEURON SURVIVAL AND NEURITE EXTENSION IN<br />

EMBRYONIC GENICULATE AND TRIGEMINAL GANGLIA<br />

Bai L. 1, Mistretta C. 1 1School of Dentistry, University of Michigan, Ann<br />

Arbor, MI<br />

Sonic hedgehog (Shh) and other Hedgehog (Hh) proteins have<br />

known roles in regulating proliferation of neuron precursors, neuron<br />

differentiation and neurite outgrowth, and Schwann cell – neuron<br />

interactions. Because Shh is intensely localized in developing papilla<br />

placodes and fungi<strong>for</strong>m papillae in embryonic rat tongue, we<br />

hypothesized that Shh, and/or other Hh proteins, also would have roles<br />

in regulating neuron number and neurite extension in the geniculate<br />

(gg) and trigeminal (tg) ganglia, which innervate these papillae. Ganglia<br />

were dissected from gestational day 13 and 16 rat embryos and<br />

explanted onto matrix-coated coverslips. Gg and tg cultures were<br />

maintained in standard medium (B27), with or without neurotrophins,<br />

BDNF or NGF, <strong>for</strong> 2 to 5 days. To test roles <strong>for</strong> Hh proteins in ganglion<br />

cultures, we added the alkaloid cyclopamine (CYCL), a specific blocker<br />

of hedgehog signaling at the receptor complex, to B27, BDNF or NGF<br />

culture media. In all cultures with CYCL, there were decreased<br />

numbers of ganglion neurons, and the extent of neurite outgrowth was<br />

reduced. CYCL effects were concentration – dependent (from 1 to 10<br />

uM). With immunohistochemistry we demonstrated that gg and tg<br />

neurons contain the hedgehog receptor proteins, patched and<br />

smoothened, both in cultures and in vivo. Another molecule in the<br />

signaling pathway, hedgehog - interacting protein, also was<br />

immunolocalized in the ganglia. The data provide evidence that<br />

members of the Hedgehog family support survival and neurite<br />

extension of embryonic gg and tg neurons. Supported by NIH NIDCD<br />

Grant 000456.<br />

53<br />

205 Poster [ ] Development of the Gustatory System<br />

GUSTATORY PHENOYPE IN DOUBLE NEUROTROPHIN<br />

KNOCKOUT MICE<br />

Nosrat I. 1, Agerman K. 2, Ern<strong>for</strong>s P. 2, Nosrat C.A. 1 1Laboratory of Oral<br />

Neurobiology, Biologic & Materials <strong>Sciences</strong>, University of Michigan,<br />

Ann Arbor, MI; 2Unit of Molecular Neurobiology, MBB, Karolinska<br />

Institutet, Stockholm, Sweden<br />

BDNF and NT-3, but not NT-4, are expressed in developing<br />

gustatory papillae and taste buds and participate in establishing the<br />

gustatory and somatosensory innervation of the tongue. BDNF, NT-3<br />

and NT-4 null mutated mice show deficits in their lingual gustatory and<br />

somatosensory innervation.<br />

We have generated double neurotrophin knockout mice and have<br />

analyzed the differences in the gustatory phenotypes of wild type,<br />

BDNF, BDNF/NT-3 and BDNF/NT-4 mice. Because BDNF/NT-3 and<br />

BDNF/NT-4 mice die at birth, we have focused our study on newborn<br />

animals. We have focused the present study on the number, average<br />

size and innervation of the gustatory papillae in these mice. Fungi<strong>for</strong>m<br />

papillae numbers were decreased significantly in all mutant mice while<br />

BDNF/NT-3 mice exhibited the most severe phenotype. Size<br />

measurements revealed a slight decrease in the size of BDNF and<br />

BDNF/NT-4 fungi<strong>for</strong>m papillae, but the strongest significant size<br />

reduction was observed in BDNF/NT-3 mice. Innervation<br />

measurements followed the same trend, with BDNF mice having the<br />

least severe loss of innervation, followed closely by BDNF/NT4 mice<br />

while BDNF/NT-3 mice exhibited the most severe phenotype. Taken<br />

together, we propose NT-3 dependent innervation plays crucial roles in<br />

maturation and maintenance of fungi<strong>for</strong>m papillae while NT-4<br />

phenotype is mimicked by BDNF in the anterior part of the tongue.<br />

206 Poster [ ] Development of the Gustatory System<br />

NEURONAL DEATH IN THE RAT GENICULATE GANGLION<br />

DURING DEVELOPMENT<br />

Carr V.M. 1, Sollars S.I. 2, Farbman A.I. 1 1Neurobiology and Physiology,<br />

Northwestern University, Evanston, IL; 2Psychology, University of<br />

Nebraska Omaha, Omaha, NE<br />

The geniculate ganglion provides sensory innervation to taste buds<br />

on the anterior tongue and the palate, and to the skin of the ear. During<br />

vertebrate development other sensory ganglia undergo a period when<br />

neuron cell death occurs, often because of insufficient target-derived<br />

trophic substance <strong>for</strong> maintenance of all ganglionic neurons. We<br />

examined neuronal death in the geniculate ganglion to try to determine<br />

when critical developmental events might occur that would be related to<br />

the survival or death of geniculate neurons. Thus far, 10 micron<br />

histological sections of geniculate ganglia from fetal (E17-E22) and<br />

newborn (P1-P3) rats have been examined following fixation in 4%<br />

para<strong>for</strong>maldehyde and standard staining procedures. Earlier stages are<br />

currently under investigation. At each age total numbers of neurons<br />

and pycnotic neurons were counted in 8 ganglia. The average incidence<br />

of pycnotic neurons at E17 was 9% of total neurons. The percentage<br />

then decreased progressively, leveling off at 0.2% by E22. Findings<br />

suggest that critical events related to neuronal survival occur from at<br />

least E17 to E21, when neuronal death is elevated. Trophic substances<br />

required <strong>for</strong> survival might be present in limited amounts so that only<br />

those neurons that receive sufficient amounts will survive. These<br />

trophic factors could originate from several sources, including the<br />

neurons themselves, the supporting cells around the neurons in the<br />

ganglion, the peripheral targets or the central targets of the neuronal<br />

projections, or a combination of sources. Supported by NIDCD<br />

Grants 04837 (AF) and 04846 (SS).


207 Poster [ ] Bitter Taste<br />

BLOCKING GLUTAMATE RECEPTORS IN THE<br />

PARABRACHIAL NUCLEUS REDUCES AVERSIVE<br />

OROMOTOR RESPONSES TO QUININE IN CONSCIOUS<br />

RATS<br />

King M.S. 1, Keller G.S. 1, Uflacker A.B. 1 1Biology, Stetson University,<br />

DeLand, FL<br />

Microinjection of glutamate into the classic gustatory region of the<br />

parabrachial nucleus (PBN), the 'waist' region (W), in conscious rats<br />

elicits ingestive oromotor behaviors, but the exact functional role of<br />

glutamate in W is unclear. This pilot study was designed to examine<br />

taste reactivity responses during oral infusion of tastants while blocking<br />

glutamate neurotransmission in W. Cannula (Plastics One) were placed<br />

bilaterally into W in 8 male Wistar rats. The cannula were connected to<br />

osmotic pumps (Alzet) that were filled with either aCSF (n=3) or<br />

kynurenate (KYN, an ionotropic glutamate receptor antagonist) in aCSF<br />

[250 (n=2), 10 (n=2) or 2.5mM (n=1)]. The pumps delivered .25µl/hr<br />

<strong>for</strong> 7 days. Rats also were implanted with intra-oral cannula <strong>for</strong> the<br />

infusion of tastants. After recovery from surgery and habituation to the<br />

behavioral arena, rats were videotaped on two successive days during 1min<br />

(.233ml) oral infusions of dH2O, .1M NaCl and .03M HCl on the<br />

first day, and dH2O, .1M sucrose and .003M quinine hydrochloride on<br />

the second day. Ingestive oromotor responses to all chemicals were<br />

similar in all groups, however aversive responses to quinine (e.g. gapes)<br />

were dramatically reduced in rats receiving infusions of 250mM KYN<br />

(mean of 56 vs 2, p<br />

+0.9) but not with responses to MgCl . These results suggest overlap in<br />

2<br />

the coding of bitter taste in<strong>for</strong>mation <strong>for</strong> the <strong>for</strong>mer three compounds<br />

but heterogeneity in the mechanisms <strong>for</strong> quinine vs. bitter salts,<br />

consistent with taste discrimination studies in rats. Additional bitter<br />

stimuli will be tested. Supported by NIH DC00353.<br />

210 Poster [ ] Bitter Taste<br />

PLC2 KNOCK-OUT MICE DISPLAY LICK AVOIDANCE TO<br />

HIGH CONCENTRATIONS OF QUININE AND DENATONIUM<br />

Dotson C.D. 1, Richter T.A. 2, Roper S.D. 2, Spector A.C. 1 1Department of<br />

Psychology and Center <strong>for</strong> Smell and Taste, University of Florida,<br />

Gainesville, FL; 2Department of Physiology and Biophysics, and<br />

Neuroscience Program, University of Miami, Miami, FL<br />

The T1R and T2R families of taste receptors apparently mediate taste<br />

receptor cell responsiveness to “sweet” and “bitter” compounds,<br />

respectively. Yet these two distinct families are believed to share key<br />

components of their signaling pathways, namely TRPM5 ion channels<br />

and an iso<strong>for</strong>m of phospholipase C (PLCβ2). There is evidence that<br />

mutant mice lacking either PLCβ2 or TRPM5 are completely<br />

unresponsive to these compounds. We sought to confirm the behavioral<br />

response characteristics of PLCβ2 knock-out (KO) mice and wild-type<br />

(WT) controls by using a brief-access taste test. Unconditioned licking<br />

responses of KO and WT mice were measured to sucrose (S), sodium<br />

chloride (N), quinine hydrochloride (Q), denatonium benzoate (D), and<br />

citric acid (C). KO mice were unresponsive to S, whereas WT mice<br />

showed concentration-dependent increases in licking. The<br />

concentration-response functions of KO mice were shifted to the right<br />

<strong>for</strong> both Q and D, but these animals nonetheless clearly showed lick<br />

suppression at higher concentrations. There were no significant<br />

differences between the C concentration-response functions from WT<br />

and KO mice. The results <strong>for</strong> N were somewhat more complex, but<br />

differences between WT and KO mice were modest. Thus mice can<br />

respond to higher concentrations of Q and D in the absence of PLCβ2<br />

suggesting that these “bitter” stimuli can activate additional<br />

transduction pathways that do not depend on this enzyme. Supported by<br />

R01-DC01628 (ACS) and R01-DC00374 (SDR).


211 Poster [ ] Bitter Taste<br />

FUNCTIONAL CHARACTERIZATION OF HUMAN T2R<br />

BITTER RECEPTORS.<br />

Sainz E. 1, Battey J.F. 1, Northup J.K. 1, Sullivan S.L. 1 1National Institute<br />

on Deafness and Other Communication Disorders, National Institutes<br />

of Health, Rockville, MD<br />

To characterize the functional properties of members of the human<br />

T2R family of bitter receptors, we are using an in vitro reconstitution<br />

assay. Baculoviral expression vectors were constructed with the human<br />

T2R genes and the mouse cycloheximide receptor gene, mT2R5.<br />

Western analyses with cellular membranes isolated from insect cells<br />

infected with these baculoviruses indicate that the membranes are<br />

highly enriched in bitter receptors. To validate our methods, we<br />

demonstrated functional reconstitution of the mT2R5 receptor by the<br />

addition of cycloheximide and purified G proteins to mT2R5-enriched<br />

membranes, as previously reported. We are currently screening orphan<br />

members of the hT2R family with a panel of 60 bitter-tasting<br />

compounds to uncover additional ligand-receptor interactions. Thus far,<br />

we have identified a bitter receptor that selectively responds to<br />

denatonium and another receptor that displays a broader response<br />

profile, responding to several of the bitter-tasting compounds tested.<br />

The pharmacophore <strong>for</strong> the seemingly more promiscuous receptor is<br />

under investigation. Given that the identities and concentrations of both<br />

the ligands and G proteins added in the assay can be controlled, this<br />

approach should allow the quantitative assessments of the ligand<br />

binding properties and the G protein selectivities of these receptors.<br />

This work is supported by the Division of Intramural Research,<br />

NIDCD/NIH/DHHS.<br />

212 Slide [ ] Bitter Taste<br />

THE EVOLUTIONARY DIVERSITY OF BITTER TASTE<br />

Hettinger T.P. 1, Frank M.E. 1 1Oral Diagnosis, Neurosciences, UCONN<br />

Health Center, Farmington, CT<br />

Bitter taste is a complex topic viewed from perspectives of chemical,<br />

receptor, behavioral and species diversity. Many chemically diverse<br />

compounds are bitter to humans, and multiple G-protein coupled<br />

receptors (GPCRs), the T2 candidate bitter receptors, have been<br />

identified. We are conducting a series of studies addressing which of<br />

those compounds may be “bitter” to golden hamsters (Mesocricetus<br />

auratus). Behavioral cross-generalizations indicate that the ionic<br />

denatonium and MgSO4, and non-ionic caffeine and sucrose<br />

octaacetate (SOA) have distinct sensory qualities to hamsters. In<br />

hamsters only the ionic stimuli activate the chorda tympani nerve (CT)<br />

and cross-generalize with quinine, the human bitter prototype. Some<br />

inbred strains of golden hamsters, e.g., ACNT, prefer nonionic SOA<br />

and caffeine, which taste bitter to humans. Hamster behavioral<br />

preference and CT neural thresholds <strong>for</strong> denatonium are 5 orders of<br />

magnitude higher than human thresholds. Thus, although denatonium<br />

may be “bitter” to hamsters and humans, the receptors are likely<br />

species-specific. Simple salts such as MgSO4 are also bitter and other<br />

mechanisms must be considered; nonetheless, candidate T2 GPCRs<br />

likely contribute to the observed bitter diversity within and across<br />

species. By analyzing nucleotide and amino-acid sequences published<br />

in the NCBI genetic database, we are working to develop models to<br />

recognize orthologous (between species) and paralogous (within<br />

species) variation among T2s in order to sort out bitter diversities.<br />

[Supported by NIH grant R01 DC04099]<br />

55<br />

213 Poster [ ] Bitter Taste<br />

HIGH RESOLUTION MAPPING OF THE BITTER TASTE<br />

SENSITIVITY LOCUS QUI<br />

Nelson T.M. 1, Munger S.D. 1, Boughter, Jr J.D. 2 1Anat & Neurobiology,<br />

Univ Maryland Sch Med, Baltimore, MD; 2Anat & Neurobiology, Univ<br />

Tennessee Hth Sci Ctr, Memphis, TN<br />

Although several genes have been implicated in the detection and<br />

transduction of taste stimuli, there is still a sizable gap in our<br />

understanding of the mechanisms of taste transduction and sensory<br />

coding, as well as how taste-related genes combine to underlie<br />

behavioral responsiveness. To better understand the genetic basis of<br />

bitter taste behavior, we have sought to identify genes that contribute to<br />

specific bitter taste sensitivities. Using a brief-access taste test, which<br />

minimizes non-gustatory variables, we determined the taste sensitivities<br />

of BXDTy recombinant inbred mouse strains (as well as their parental<br />

strains, C57BL/6J (B6) and DBA/2J (D2)) to several bitter stimuli. Our<br />

initial interval mapping of the BXDTy dataset confirmed a significant<br />

quantitative trait locus (QTL), Qui, <strong>for</strong> quinine taste sensitivity (e.g.,<br />

Lush, 1986) on distal Chr 6. We have defined the physical interval that<br />

contains Qui. This interval includes a cluster of 24 Tas2r genes<br />

encoding T2R putative bitter taste receptors. Comparisons of the<br />

protein coding sequences across B6, D2 and BXDTy lines revealed<br />

distinct B6 and D2 alleles <strong>for</strong> a number of Tas2r genes; these allelic<br />

variants correlate with the taster status of individual BXDTy lines.<br />

These studies should facilitate the identification of major quantitative<br />

trait genes underlying bitter taste. Supported by NIDCD:<br />

DC005786(SDM), DC004935(JDB).<br />

214 Poster [ ] Bitter Taste<br />

RELATIONSHIP BETWEEN GENOTYPES OF THE TAS2R38<br />

GENE AND BITTER PERCEPTION IN<br />

Mennella J.A. 1, Pepino M.Y. 1, Kennedy J.M. 1, Mascioli K.J. 1, Reed<br />

D.R. 1 1Monell Chemical Senses Center, Philadelphia, PA<br />

The present study aimed to determine how variation in the TAS2R38<br />

gene influences the gustatory experience and preferences of children<br />

and adults <strong>for</strong> bitterness. To this aim, genotype and behavioral analyses<br />

were per<strong>for</strong>med on 5- to 10-year-old children of African (43 girls, 44<br />

boys) or European (31 girls, 23 boys) descent and their mothers.<br />

Genomic DNA was extracted from cheek cells and alleles of the<br />

TAS2R38 gene were genotyped <strong>for</strong> the variant sites using allelespecific<br />

probes and primers. Forced-choice procedures embedded in the<br />

context of a game were used to determine sensitivity to the bitter taste<br />

of 6-n-propylthiouracil (PROP) during one test session and food habits<br />

during another. Regardless of race, alleles at TAS2R38 were<br />

significantly associated with PROP thresholds in children (p


215 Poster [ ] Bitter Taste<br />

BITTER TASTE MARKERS IDENTIFY SWEET AND<br />

ALCOHOL HEDONICS AND INTAKE<br />

Dinehart M.E. 1, Bartoshuk L. 2, Kinsley E. 1, Duffy V.B. 1 1Dietetics,<br />

University of Connecticut, Storrs, CT; 2Surgery, Yale University, New<br />

Haven, CT<br />

Human and animal data suggest connections between sweet and<br />

alcohol behaviors. We tested the ability of markers of variation in taste<br />

to identify preference <strong>for</strong> and intake of sweet foods and alcoholic<br />

drinks. Following the pioneering work of Fischer et al from the 1960s,<br />

taste variation was characterized by bitterness of 6-n-propylthiouracil<br />

(PROP) and quinine hydrochloride (QHCl). Bitterness varies<br />

genetically and across the lifespan (eg, hormonally, exposure to taste<br />

pathology). Adults (74 F, 51 M; 22-58 years) used the general Labeled<br />

Magnitude Scale to rate bitterness of 3.2 mM PROP and .32 mM QHCl<br />

and preference <strong>for</strong> sampled and survey sweet foods. Subjects reported<br />

alcohol and sweet food intake via frequency interview and energy from<br />

added sugar via food records analyzed with the USDA Pyramid<br />

Servings Database. Multiple regression revealed that lower preference<br />

<strong>for</strong> or intake of sweets and alcohol was predicted by greater PROP<br />

bitterness but lower QHCl bitterness. Although PROP and QHCl<br />

showed significant correlation, subjects fell into groups that were<br />

discordant in these bitters but nearly matched <strong>for</strong> sex. Those who tasted<br />

QHCl as more bitter relative to PROP (n=23) had significantly greater<br />

preference <strong>for</strong> and intake of sweets and greater alcohol intake than<br />

those who taste PROP as more bitter relative to QHCl (n=24). In<br />

summary, use of both bitter makers increased the ability to explain<br />

sweet and alcohol behaviors. Whether quinine bitterness reflects<br />

another genetic taste variant or an environmentally-mediated pattern of<br />

sensation is unknown. (NRICGP/USDA 2002-00788, NIH DC00283)<br />

216 Poster [ ] Olfactory Bulb: Coding<br />

INHIBITION OF ADENYLYL CYCLASE IN LOBSTER<br />

OLFACTORY RECEPTOR NEURONS ENHANCES CENTRAL<br />

RESPONSES TO ODORS<br />

Aggio J.F. 1, Daly K. 2, Ache B. 1 1The Whitney Laboratory, University of<br />

Florida, St. Augustine, FL; 2Dept. of Entomology, Ohio State<br />

University, Columbus, OH<br />

Olfactory receptor neurons (ORNs) in some animals can be inhibited<br />

as well as excited by odorants. The question is whether such bipolar<br />

input is functionally significant in olfaction. Pharmacological<br />

intervention of the cyclic nucleotide signaling cascade in lobster ORNs<br />

selectively alters inhibitory responses to odorants. As a first step to<br />

addressing this question, we assessed the effect of blocking adenylyl<br />

cyclase (AC) peripherally on the output of the first olfactory relay, the<br />

olfactory lobe (OL), in a perfused lobster nose-brain preparation. We<br />

show that pharmacologically blocking AC in the periphery had no<br />

effect by itself, but could enhance the response of the OL to odorant<br />

mixtures, consistent with the predicted reduction in odorant-evoked<br />

inhibitory input that should follow AC blockade. Blocking AC could<br />

also enhance the response of the OL to single component odorants.<br />

Given earlier evidence that monomolecular odorants excite and inhibit<br />

different cells, the latter finding suggests that spontaneous activity in<br />

the ORNs may be sufficient to register peripheral inhibition in the OL<br />

in the absence of co-activation by other odorants. Our results suggest<br />

that inhibitory input to the olfactory periphery helps shape the output of<br />

the OL and, there<strong>for</strong>e, potentially contributes to the olfactory code.<br />

Supported by the McDonnell Foundation (BWA, KCD) and the<br />

NIDCD (DC05535, KCD).<br />

56<br />

217 Poster [ ] Olfactory Bulb: Coding<br />

INTER- AND INTRA-SPECIES ANTENNAL IMAGINAL DISC<br />

TRANSPLANTS: BEHAVIOR, SENSORY AND CENTRAL<br />

OLFACTORY NEUROPHYSIOLOGY<br />

Hillier K.N. 1, Vickers N.J. 1, Linn C. 2 1Biology, University of Utah, Salt<br />

Lake City, UT; 2Entomology, Cornell University, Geneva, NY<br />

Vertebrate and invertebrate organisms share a common feature in<br />

having the primary olfactory neuropil organized into discrete knots<br />

known as glomeruli. The macroglomerular complex (MGC) is a<br />

sexually dimorphic cluster of male-specific, pheromone-receptive<br />

glomeruli that is configured differently between moth species. The<br />

physiology and arrangement of MGC glomeruli is dictated, in part, by<br />

the antennal imaginal disc. We have previously demonstrated that premetamorphic<br />

inter-species transplantation of antennal imaginal discs<br />

resulted in the development of a 'donor' type MGC. Our current study<br />

seeks to gain further insight into the factors affecting changes in<br />

olfactory physiology and behavior through unilateral transplantation of:<br />

a) a Heliothis virescens disc into Helicoverpa zea (V-Zr), and b) a H.<br />

virescens disc into a H. virescens recipient (V-Vr). To test moth<br />

behavioral discrimination, unoperated antennae were removed and the<br />

ability of moths to differentiate between each species' pheromone<br />

blends was tested in a wind tunnel. Both transplant types had similar,<br />

unusual sensory neurons tuned to Z9-16:Ald, an essential component of<br />

the normal H. zea pheromone blend. However, only V-Zr males<br />

responded behaviorally to blends containing this compound. Cobaltlysine<br />

stains of the sensilla housing these unexpected sensory neurons<br />

and recordings from central interneurons suggested that this behavioral<br />

difference was the result of the specific MGC glomeruli activated by<br />

this compound. Supported by NIH-1 R55 DC04443-01 to CEL<br />

218 Poster [ ] Olfactory Bulb: Coding<br />

MACROGLOMERULI IN THE WORKER CASTE OF LEAF-<br />

CUTTING ANTS<br />

Kleineidam C.J. 1, Obermayer M. 1, Halbich W. 1, Roessler W. 1<br />

1University of Wuerzburg, Wuerzburg, Germany<br />

Workers of leaf-cutting ants express an extraordinary size<br />

polymorphism with a factor of 1000 in body mass. Foraging workers<br />

show less variation but still can be discriminated in minor and medium<br />

workers. We asked the question whether these workers differ in a) their<br />

morphology of the first olfactory neuropil, the antennal lobe, b) their<br />

olfactorial behavior to trail pheromones and c) their receptor neuron<br />

response to trail pheromones. We found, <strong>for</strong> the first time in nonsexual<br />

individuals, a greatly enlarged glomerulus. The comparison of two<br />

closely related species Atta sexdens and Atta vollenweideri by 3Dreconstructions<br />

of the antennal lobes revealed striking similarities as<br />

well as very distinct differences in the arrangement of macroglomeruli<br />

among the two species. Size polymorphism is found in antennal lobe<br />

structures, the glomeruli. While medium workers have a<br />

macroglomerulus, the glomeruli of the minor workers are all of similar<br />

size. We tested the antenna in EAGs with two common and main<br />

components of the trail pheromone of leaf-cutting ants (4-Methylpyrrol-<br />

2-Carboxylat and 2-Ethyl-3,5-Dimethylpyrazine) and found that the<br />

relative response to those two components differs significantly. If a<br />

larger glomerulus reflects a larger number of terminating receptor<br />

neurons, this result supports the idea that the macroglomerulus is<br />

involved in trail detection. In behavioral tests we found that trail<br />

following behavior is somewhat lower in minor workers than in<br />

medium workers. Surprisingly, in a choice experiment with gland<br />

extracts of nestmates and of the comparison species the minor workers<br />

outper<strong>for</strong>med the medium workers. Funding: DFG; SFB 554


219 Poster [ ] Olfactory Bulb: Coding<br />

THE EFFECTS OF STIMULUS DYNAMICS ON OLFACTORY<br />

LOBE RESPONSES IN THE CRAYFISH, PROCAMBARUS<br />

CLARKII USING ENSEMBLE RECORDING TECHNIQUES<br />

Wolf M. 1, Daly K. 2, Moore P.A. 1 1Biological <strong>Sciences</strong>, Bowling Green<br />

State University, Bowling Green, OH; 2Entomology, Ohio State<br />

University, Columbus, OH<br />

Two major issues facing sensory ecologists are how olfactory<br />

in<strong>for</strong>mation is coded in the nervous system and that in<strong>for</strong>mation<br />

correlates with behavior. Behavioral results from our lab indicate that<br />

crayfish are sensitive and respond to temporal components of an odor<br />

signal. As such, we expect to find that underlying this sensitivity to<br />

stimulus dynamics, are neural correlates evident within the olfactory<br />

system. Thus, in this preliminary study, we investigated how the<br />

temporal aspects of an odor signal are coded in the olfactory lobe of the<br />

crayfish. Neural ensemble recordings were made on an isolated head<br />

preparation perfused with oxygenated crayfish saline. Silicon<br />

multichannel electrode arrays were inserted into the olfactory lobe of<br />

the crayfish brain. The medial antennule was placed into an<br />

olfactometer and stimulated with 3 types of stimuli; glutamate, glycine,<br />

and shrimp extract. The stimuli were presented at a specific molar<br />

concentration (10-5 M) and duration (500 ms), varying only the<br />

intermittency between odor pulses. Our results suggest that coordinated<br />

clusters of units, which collectively produced odor-dependent responses<br />

but these responses were further dependent on stimulus intermittency.<br />

These results are consistent with our behavioral data demonstrating that<br />

crayfish are sensitive to the manner in which odors are experienced.<br />

This work was supported by the McDonnell Foundation (KCD),<br />

NIDCD-DC05535, KCD and a Sigma Xi Grant-in-aid of research<br />

(MCW)<br />

57<br />

220 Poster [ ] Olfactory Bulb: Coding<br />

THE EFFECT OF STIMULUS DURATION ON EUCLIDIAN<br />

RESPONSE DISTANCE MEASURES OF ODOR<br />

DISCRIMINATION ACROSS ANTENNAL LOBE<br />

POPULATIONS IN MANDUCA SEXTA.<br />

Daly K.C. 1, Smith B.H. 1 1Entomology, Ohio State University,<br />

Columbus, OH<br />

Behavioral studies suggest that animals such as the moth Manduca<br />

sexta perceive subtle differences between odorant based on changes<br />

molecular features such a carbon chain length. Recent theories of<br />

temporal coding predict that closely related odors are “decomposed”<br />

over time and that longer duration stimulations should provide<br />

additional in<strong>for</strong>mation used by the antennal lobe (AL) or olfactory bulb<br />

to enhance discrimination. To test this we measured the degree to which<br />

AL ensembles could statistically discriminate closely related odors as a<br />

function of increasing stimulus duration. Multiunit recordings from<br />

moths ALs were analyzed in response to 6 alcohols and ketones that<br />

were repeatedly presented at odor pulse durations ranging from 50 to<br />

4000 ms. Principle Components Analysis, on binned data (10ms),<br />

extracted orthogonal factors each representing a collection of units with<br />

common and coordinated responses. General Linear Modeling<br />

identified factors with odor-dependent temporal effects (p< 0.001).<br />

Odor-dependent factors were treated as independent dimensions and the<br />

Euclidian distance between odor responses at bin was calculated in a<br />

high dimensional space. We find that population trajectories rapidly<br />

peak (animal specific ~120-240 ms) and does so at the same time<br />

irrespective of stimulus duration. For longer durations, trajectories do<br />

tend to stay separated <strong>for</strong> the duration of the pulse length. These<br />

preliminary results suggest that olfactory systems have the capacity to<br />

represent odorant identity of subtly different odors rapidly. Supported<br />

by NIH-NCRR; RR14166-06 (BS) & NIH-NIDCD; DC05535-01 and<br />

the McDonnell Foundation (KD)<br />

221 Poster [ ] Olfactory Bulb: Coding<br />

CHARACTERIZATION OF LABELED CELLS IN THE<br />

OLFACTORY BULB OF TRANSGENIC ZEBRAFISH<br />

EXPRESSING THE SIMIAN CYTOMEGALOVIRUS (SCMV)<br />

PROMOTER<br />

Fuller C.L. 1, Suhr S.T. 2, Goldman D.J. 2, Byrd C.A. 1 1Biological<br />

<strong>Sciences</strong>, Western Michigan University, Kalamazoo, MI; 2Mental<br />

Health Research Institute, University of Michigan, Ann Arbor, MI<br />

The CMV promoter is commonly used <strong>for</strong> the production of<br />

transgenic animals due to its effective and essentially ubiquitous<br />

expression. We used a simian CMV promoter to drive the expression of<br />

a fluorescent reporter protein, dsRed, in zebrafish. Expression of the<br />

sCMV promoter has been confirmed in several regions throughout the<br />

nervous system including a specific subset of cells in the olfactory bulb.<br />

These olfactory bulb cells were among the first to express the transgene<br />

early in development. The purpose of this study was to examine the<br />

morphology, distribution, and identity of the labeled cells in the<br />

olfactory bulb using confocal microscopy, immunocytochemistry, and<br />

retrograde tract-tracing methods. The labeled cells have teardropshaped<br />

somata that are 5-10 microns in diameter. The cell bodies are<br />

found throughout the glomerular and superficial internal cell layers, and<br />

they possess a single prominent process containing tufts in the<br />

glomerular layer. Although these cells are found in the same location as<br />

mitral cells, they do not appear to be output neurons since retrograde<br />

labeling of the olfactory tracts with a fluorescent dextran identifies a<br />

different subset of bulbar cells. We are continuing our exploration of<br />

the identity of the bulbar neurons that are labeled in these transgenic<br />

zebrafish.<br />

Supported by NIH DC04262 (CAB) and a grant from the Hereditary<br />

Disease Foundation: Cure Huntington´s Disease Initiative (STS).


222 Poster [ ] Olfactory Bulb: Coding<br />

CADHERIN AND CATENIN EXPRESSION IN THE<br />

OLFACTORY NERVE<br />

Akins M. 1, Greer C.A. 2 1Neurosurgery, Yale University, New Haven,<br />

CT; 2Neurobiology, Yale University, New Haven, CT<br />

Olfactory sensory neurons extend axons to the olfactory bulb, where<br />

they <strong>for</strong>m the olfactory nerve layer (ONL), which is subdivided into<br />

outer (ONLo) and inner (ONLi) sublaminae. In the ONLo axons<br />

primarily undergo extension and gross targeting while targeting specific<br />

glomeruli occurs within the ONLi. Mechanisms <strong>for</strong> differential axoaxonal<br />

adhesion within the ONLi versus the ONLo are not known.<br />

However, the cadherin family of molecules are candidates. Together<br />

with their intracellular binding partners, the catenins, they mediate<br />

homotypic cell-cell adhesion and have been implicated in many roles<br />

during neural development, including axonal extension and specific<br />

circuit <strong>for</strong>mation. Using immunohistochemistry, we have localized Ncadherin<br />

(CDH2) and several catenins in the mouse olfactory pathway.<br />

CDH2 and several catenins are more strongly expressed in the ONLo<br />

than the ONLi during perinatal development, a pattern that is absent in<br />

adulthood. γ-Catenin, which binds directly to cadherins and links to the<br />

cytoskeleton, is expressed strongly in the ONLo, as is δ-catenin, which<br />

binds directly to cadherins and modulates their function. In contrast βcatenin,<br />

which competes with γ-catenin <strong>for</strong> cadherin binding, is<br />

expressed uni<strong>for</strong>mly in the ONLo and ONLi, as is p120-catenin, which<br />

competes with δ-catenin. These findings suggest the presence of two<br />

different cadherin-based adhesion systems in the ONL, one of which<br />

contains CDH2, γ-catenin, and δ-catenin and helps restrict axons to the<br />

ONLo during extension. Preliminary evidence further suggests that<br />

intermediate filaments may be involved in the function of this adhesion<br />

system.<br />

DC006335, DC00210<br />

223 Poster [ ] Olfactory Bulb: Coding<br />

ACTION POTENTIAL BACKPROPAGATION AND MODULAR<br />

PROCESSING OF VOMERONASAL RECEPTOR INPUT IN<br />

RAT ACCESSORY OLFACTORY BULB<br />

Ma J. 1, Lowe G. 1 1Monell Chemical Senses Center, Philadelphia, PA<br />

In main olfactory bulb, receptor cells expressing one OR gene project<br />

to unique paired glomeruli. Mitral cells link to single glomeruli via<br />

simple primary dendrites which reliably propagate action potentials.<br />

Accessory olfactory bulb (AOB) exhibits more complex wiring:<br />

sensory neurons (VSNs) expressing one VR gene make divergent<br />

projections to many glomeruli, and mitral cells link to multiple<br />

glomeruli via branched primary dendrites. We applied Ca-imaging to<br />

track backpropagating action potentials (BPAPs) in rat AOB primary<br />

dendrites. Under whole-cell dialysis with 100 µM Ca-orange, somatic<br />

spike-evoked fluorescent signals were detected over the entire dendritic<br />

tree; ∆F/F was nearly constant on all branches from soma to glomeruli.<br />

Backpropagation relied on Na + channels: in 1 µM TTX, somatic AP<br />

voltage-clamp commands evoked dendritic Ca-transients that declined<br />

significantly with distance compared to current-clamp controls. Dual<br />

soma-dendrite recording confirmed that BPAPs were unattenuated,<br />

while subthreshold voltage transients declined markedly. Ca-transients<br />

were not significantly altered by 100 µM APV or 50 µM bicuculline,<br />

suggesting BPAPs are unaffected by local synaptic input. Genetic<br />

tracing in AOB has suggested homotypic connectivity - individual<br />

dendritic arbors project only to glomeruli targeted by VSNs expressing<br />

the same VR gene. Non-decremental, non-dichotomous<br />

backpropagation in AOB primary dendrites ensures fast, reliable<br />

communication between mitral cells and their homotypic glomeruli,<br />

binding them into coherent modules in accordance with their VR-coded<br />

inputs. Supported by NIDCD DC04208 (GL).<br />

58<br />

224 Poster [ ] Olfactory Bulb: Coding<br />

MITRAL/TUFTED AND GRANULE CELL RESPONSE<br />

SPECIFICITY IN THE MOUSE OLFACTORY BULB.<br />

Davison I.G. 1, Shtoyerman E. 1, Katz L.C. 1 1Dept. of Neurobiology,<br />

Duke University Medical Center, Durham, NC<br />

Odor representations passed from the olfactory bulb to higher brain<br />

centers are contained in the firing patterns of olfactory bulb<br />

mitral/tufted (M/T) neurons. The degree of stimulus specificity of M/T<br />

neurons has important implications <strong>for</strong> distinguishing between<br />

competing hypotheses of olfactory coding. Assessing stimulus<br />

specificity has been difficult due to the huge size of `odor space´: the<br />

range of potentially active volatile chemicals that could activate a<br />

neuron. To examine M/T response specificity, we use a computercontrolled,<br />

robotic odor delivery system that allows fast and simple<br />

application of hundreds of odorants with precisely controlled dilution<br />

and timing. Extracellular single unit recordings from the mouse OB in<br />

vivo, using odorant concentrations within a behaviorally relevant range,<br />

revealed that M/T responses are extremely specific. When presented<br />

with a large panel of structurally diverse and closely related categories,<br />

M/T neurons typically respond to


226 Poster [ ] Olfactory Bulb: Coding<br />

ONTOGENY OF ODOR DISCRIMINATION<br />

Fletcher M. 1, Wilson D. 1 1Zoology, University of Oklahoma, Norman,<br />

OK<br />

Individual olfactory bulb mitral/tufted cells respond preferentially to<br />

groups of molecularly similar odorants. Bulbar interneurons are<br />

thought to influence mitral/tufted odorant receptive fields (RFs) through<br />

mechanisms such as lateral inhibition. Infant rats, however, lack a<br />

majority of these inhibitory interneurons until the second week of life.<br />

It is unclear if these developmental differences affect mitral/tufted RFs<br />

or affect behavioral odor discrimination. The following experiments<br />

aimed at better understanding olfactory bulb mitral/tufted cell receptive<br />

fields, odor coding, and behavioral odor discrimination in the<br />

developing olfactory system.<br />

Single-unit and local field potential recordings were made from<br />

mitral/tufted cells of freely breathing urethane-anesthetized rats (PN4 –<br />

adult). RFs to a homologous series of esters of different carbon chain<br />

lengths were mapped <strong>for</strong> each age group. Odorants were equated <strong>for</strong><br />

concentration (150 PPM) using a flow dilution olfactometer.<br />

Preliminary results suggest minimal differences between infant and<br />

adult single-unit mitral/tufted RFs. However, odor-evoked local field<br />

potential oscillations showed a strong age-dependent change in<br />

dominant frequency over the age range tested. Behavioral odor<br />

discrimination to the same set of odorants is currently being examined<br />

using our cardiac orienting response paradigm, which is effective even<br />

in very young (PN4) animals.<br />

Supported by: DC03906 to DAW and F31 DC006126 to MLF<br />

227 Poster [ ] Olfactory Bulb: Coding<br />

EFFECTS OF FUNCTIONAL GROUP POSITION ON<br />

GLOMERULAR ACTIVATION PATTERNS EVOKED BY<br />

ESTER AND ALCOHOL ODORANTS.<br />

Johnson B.A. 1, Leon M. 1 1Neurobiology and Behavior, University of<br />

Cali<strong>for</strong>nia, Irvine, Irvine, CA<br />

Studies of the relationship between odorant structure and bulbar<br />

activity patterns have identified glomerular modules that recognize<br />

discrete odorant molecular features such as functional groups. In the<br />

present experiments, we asked if the position of functional groups<br />

within odorants influences spatial activity patterns. By mapping uptake<br />

of [14C]2-deoxyglucose across the entire glomerular layer, we<br />

compared the influences of functional group position and carbon<br />

number using systematically differing series of 21 alcohols and 16<br />

esters. Along every carbon number and positional series,<br />

representations were chemotopic, with the most similar chemicals<br />

evoking the most similar patterns. However, the relative impact of the<br />

position of the functional group differed greatly <strong>for</strong> esters and alcohols.<br />

Changing the position of the ester functional group determined whether<br />

or not particular glomerular modules were activated, but caused only<br />

small differences in the overall glomerular activity pattern. In contrast,<br />

changing the position of the alcohol functional group evoked very<br />

distinct patterns involving entirely different groups of glomerular<br />

modules. Thus, the position of functional groups is an important feature<br />

determining patterns of glomerular activity, although the nature of the<br />

difference in the evoked response depends on the functional group.<br />

Supported by grant DC03545.<br />

59<br />

228 Poster [ ] Olfactory Bulb: Coding<br />

RESPONSES TO KETONES ARE NOT ORGANIZED<br />

CHEMOTOPICALLY WITHIN A KETONE-RESPONSIVE<br />

GLOMERULAR MODULE.<br />

Farahbod H. 1, Johnson B.A. 1, Leon M. 1 1Neurobiology and Behavior,<br />

University of Cali<strong>for</strong>nia, Irvine, Irvine, CA<br />

We have shown previously that every odorant activates a unique<br />

combination of glomerular modules in the rat olfactory bulb. Some of<br />

the modules appear to respond to a range of odorant molecules sharing<br />

a chemical feature such as a functional group. Within modules<br />

responding to aldehydes, alcohols, and acids, responses are arranged<br />

chemotopically with respect to molecular length. Ketones activate a<br />

module located dorsomedially in the bulb and in the present study, we<br />

investigated the possibility of a chemotopic organization within this<br />

module. We mapped uptake of radiolabeled 2-deoxyglucose in the<br />

olfactory bulbs of rats during stimulation by straight-chain ketone<br />

odorants that differed systematically in their carbon numbers (2pentanone<br />

to 2-undecanone) or in the position of their carbonyl group<br />

(2- and 3-hexanone and 2-, 3-, and 4-heptanone). As the carbon number<br />

increased, the pattern of activity in the posteriolateral, and medial parts<br />

of the bulb shifted toward ventral positions. However, unlike the<br />

responses of other functional groups, there was no change in the<br />

location of the responses within the ketone-responsive module as<br />

carbon number increased. Changing the position of other functional<br />

groups has been shown to affect odorant response patterns, but such<br />

changes were not seen within the ketone-responsive module. These data<br />

indicate that the organization of responses within the olfactory system<br />

differs across odorant functional groups. Supported by grant DC03545.<br />

229 Poster [ ] Olfactory Bulb: Coding<br />

INFORMATICS TOOLS FOR GLOBAL MAPPING OF ODOR-<br />

INDUCED NEURAL ACTIVITY IN THE GLOMERULAR<br />

LAYER OF THE RODENT OLFACTORY BULB<br />

Liu N. 1, Xu F. 2, Shepherd G.M. 2, Miller P.L. 1 1Center <strong>for</strong> Medical<br />

In<strong>for</strong>matics, Yale University, New Haven, CT; 2Neurobiology, Yale<br />

University, New Haven, CT<br />

Aims: We are developing computer software tools, OdorMapBuilder<br />

and OdorMapComparer, to generate and analyze odor maps in the<br />

rodent olfactory system. Methods: The software programs are written in<br />

the Java programming language. OdorMapBuilder allows users to trace<br />

the glomerular layer in a MRI slab, which then serves as a template <strong>for</strong><br />

the glomerular fMRI density slab. The patterns from all slabs are<br />

combined into a 2D map image. Each position in the odor map<br />

represents a unique site in the OB glomerular layer, with its optical<br />

density representing the neural activity. To aid in analyzing the maps,<br />

OdorMapComparer allows users to import two maps onto a framed<br />

canvas, per<strong>for</strong>m warping, and carry out simple addition, subtraction and<br />

statistical analysis between the two images. Results: The programs<br />

provide a user-friendly interface with a rich set of menus.<br />

OdorMapBuilder generates odor maps in different perspectives: dorsal,<br />

lateral, ventral and medial. The map images can be saved in JPEG and<br />

GIF <strong>for</strong>mat or exported as raw pixel data. OdorMapComparer enables<br />

quantitative comparisons of two images and calculations of the<br />

difference or additive effects of any two odor maps. Conclusions: The<br />

present study illustrates the critical role of in<strong>for</strong>matics tools in<br />

analyzing the neural basis of the olfactory processing. In addition to<br />

fMRI data, these tools also apply to 3D data obtained using other<br />

methods. Acknowledgments: Supported by the Human Brain Project<br />

and the National Library of Medicine.


230 Poster [ ] Olfactory Bulb: Coding<br />

LATERAL INHIBITION: IT MAKES SCENTS AS A<br />

NEURONAL CODING STRATEGY IN OLFACTION<br />

Lei H. 1, Reisenman C. 1, Christensen T.A. 1, Hildebrand J.G. 1 1ARL Div.<br />

of Neurobiology, University of Arizona, Tucson, AZ<br />

There is good evidence that the primary representation of an<br />

olfactory stimulus is shaped by inhibitory interactions between<br />

glomeruli, but there are few olfactory systems in which the odor-tuning<br />

properties of identifiable and accessible glomeruli are known. The moth<br />

antennal lobe contains a small number of sexually dimorphic glomeruli:<br />

the macroglomerular complex (MGC) in males and the large female<br />

glomeruli (LFGs) in females, and evidence is increasing that the MGC,<br />

LFGs, and the remaining glomeruli share similar principles of synaptic<br />

organization. To test this idea further, we used intracellular recording<br />

and staining to examine the odor-evoked responses of projection<br />

neurons (PNs) innervating spatially identifiable glomeruli in both sexes.<br />

In males, our earlier results showed that the synchronous firing of PNs<br />

innervating one MGC glomerulus is modulated by local inhibition from<br />

the neighboring glomerulus. Similarly, in females, we found that PNs<br />

innervating one identifiable glomerulus (G40) were depolarized<br />

selectively by cis-3-hexenyl acetate, while PNs innervating the lateral<br />

LFG (which neighbors G40) were activated selectively by (+)linalool.<br />

In contrast, (±)linalool evoked a rapid inhibitory potential in PNs<br />

innervating the adjacent G40 glomerulus. Similarly, (+) but not (-)<br />

linalool also evoked an IPSP in medial LFG-PNs. These results provide<br />

direct evidence <strong>for</strong> odor-mediated, lateral inhibitory interactions<br />

between glomeruli in both the pheromonal and non-pheromonal<br />

olfactory subsystems in this insect. Pharmacological tests that<br />

selectively influence GABAergic transmission are expected to shed<br />

light on the neurochemical basis <strong>for</strong> these inhibitory synaptic<br />

interactions.<br />

Supported by grants from the NIH/NIDCD<br />

231 Poster [ ] Olfactory Bulb: Coding<br />

CONFIGURATIONAL AND ELEMENTAL ODOR MIXTURE<br />

PERCEPTION CAN ARISE FROM LOCAL INHIBITION<br />

Cleland T. 1, Linster C. 1 1Cornell University*, Ithaca, NY<br />

Contrast enhancement via lateral inhibitory circuits is a common<br />

mechanism in sensory systems. We here employ a computational model<br />

to show that, in addition to shaping experimentally observed molecular<br />

receptive fields in the olfactory bulb, functionally lateral inhibitory<br />

circuits can also mediate the elemental and configurational properties of<br />

odor mixture perception. To the extent that odor perception can be<br />

predicted by slow-timescale neural activation patterns in the olfactory<br />

bulb, and to the extent that interglomerular inhibitory projections map<br />

onto a space of odorant similarity, the model shows that these inhibitory<br />

processes in the olfactory bulb suffice to generate the behaviorally<br />

observed inverse relationship between two odorants´ perceptual<br />

similarities and the perceptual similarities between either of these same<br />

odorants and their binary mixture.<br />

60<br />

232 Slide [ ] Olfactory Bulb: Coding<br />

HIGH-DIMENSIONAL CONTRAST ENHANCEMENT IN<br />

ODOR SPACE<br />

Cleland T.A. 1, Sethupathy P. 1 1Neurobiology & Behavior, Cornell<br />

University*, Ithaca, NY<br />

The topographical mapping of external stimulus spaces is a common<br />

principle of neural organization, and contributes to sensory input<br />

processing by facilitating contrast enhancement mechanisms such as<br />

center-surround lateral inhibition. This process requires that the<br />

intrinsic topology of the contrast enhancement mechanism match the<br />

underlying topology of the sensory space upon which it acts. For<br />

example, the spatial contrast of retinal images is enhanced by lateral<br />

inhibitory projections along the two dimensions of the retinal field, and<br />

auditory frequency tuning in the inferior colliculus is similarly<br />

sharpened by lateral inhibition along the single dimension of frequency<br />

space.<br />

Contrast enhancement in the olfactory system is believed to<br />

sharpen odor quality representations, which map onto an indeterminate<br />

but certainly high-dimensional sensory space. This high-dimensional<br />

odor space must map onto functionally two-dimensional neural cortices<br />

while retaining unambiguous high-dimensional similarity relationships<br />

that are common among conspecifics; hence, physical proximity cannot<br />

reliably connote stimulus similarity. For a "lateral-inhibitory"<br />

mechanism of contrast enhancement to function, bulbar activation by a<br />

given odorant feature must specifically inhibit numerous other bulbar<br />

regions activated by chemically similar odorant features and sparsely<br />

distributed in space; current models have not addressed the scope of this<br />

fundamentally high-dimensional problem. We here propose and<br />

demonstrate a novel bulbar mechanism <strong>for</strong> such hyperlateral inhibition<br />

in odor similarity space, and challenge this model with evidence from<br />

the experimental literature.<br />

233 Poster [ ] Olfactory Bulb: Coding<br />

GLOMERULAR ON-OFF MODEL OF OLFACTORY CODING<br />

Rinberg D. 1, Gelperin A. 1, Koulakov A. 2 1Monell chemical Senses<br />

Center, Philadelphia, PA; 2Cold Spring Harbor Laboratory, Cold<br />

Spring Harbor, NY<br />

We present a model <strong>for</strong> olfactory coding based on spatial<br />

representation of glomerular responses. In this model distinct odorants<br />

activate specific subsets of glomeruli, dependent upon the odorant´s<br />

concentration. The glomerular response specificities are understood<br />

statistically, based on experimentally measured distributions of odor<br />

detection thresholds. A simple version of the model, in which<br />

glomerular responses are binary (the on-off model), allows us to<br />

quantitatively account <strong>for</strong> the following results of human/rodent<br />

psychophysics: 1) just noticeable differences in perceived concentration<br />

of a single odor (Weber ratios) are dC/C ~ 0.1; 2) the number of<br />

simultaneously perceived odors can be as high as 12 (Jinks & Laing,<br />

1999); 3) extensive lesions of the olfactory bulb do not lead to<br />

significant changes in detection or discrimination thresholds. A more<br />

detailed model allows us to reproduce closely the conditional<br />

probabilities obtained in human psychophysical experiments on<br />

perception of complex odors.


234 Poster [ ] Social Odors and Behavior<br />

UNDERSTANDING CHEMICAL COMMUNICATION UNDER<br />

LOTIC AND LENTIC CONDITIONS IN THE LABORATORY<br />

WITH CRAYFISH<br />

Redman C. 1, Bergman D.A. 1, Moore P.A. 1 1Biological <strong>Sciences</strong>,<br />

Bowling Green State University, Bowling Green, OH<br />

The physics of environments can structure how animals send and<br />

receive signals. In fact, habitat specific physics may constrain signal<br />

transmission and provide a mechanism <strong>for</strong> evolutionary sensory biases.<br />

This experiment investigates how the use of chemical (olfactory)<br />

signals to convey social signals is influenced by environmental physics.<br />

If environments can place constraints upon chemical communication,<br />

then we would predict that crayfish found in lotic (flowing water)<br />

systems should be adapted <strong>for</strong> more effective communication within<br />

this environment and crayfish in lentic (low or no flow) systems should<br />

be adapted <strong>for</strong> effective communication in this environment. This<br />

hypothesis extends the “Matched Filter” idea of Wehner to the level of<br />

behavior. We hypothesize that crayfish are influenced by the<br />

environment, thus dominant crayfish collected from lotic systems will<br />

take residence upstream in the flow, whereas under lentic (no flow)<br />

conditions the position of dominant individuals will be random. Lotic<br />

conditions consisted of two flow regimes (5 cm/s and 10 cm/s) and<br />

lentic conditions had no flow (0 cm/s). A second experiment<br />

demonstrated when urine was released under these conditions. The<br />

crayfish received an injection of 0.1% sodium Fluorescein at a dose of<br />

2-6 ug g-1 body mass. The Fluorescein injection results in the<br />

visualization of urine released during the course of an agonistic bout.<br />

This section of the study examined the ability to project urine in the<br />

different flow regimes. This study is a model <strong>for</strong> elucidating how<br />

environments structure communication systems and also whether<br />

crayfish release urine in manner that suggests that the signal is<br />

deliberately released to communicate status.<br />

235 Poster [ ] Social Odors and Behavior<br />

HPLC ANALYSIS OF THE CHEMICAL COMPOSITION OF<br />

URINE IN THE CRAYFISH, ORCONECTES RUSTICUS<br />

Martin A. 1, Bergman D. 1, Moore P.A. 1 1Biological <strong>Sciences</strong>, Bowling<br />

Green State University, Bowling Green, OH<br />

Social communication is important <strong>for</strong> the <strong>for</strong>mation of social<br />

hierarchies in crayfish. Communication uses various sensory modalities<br />

that play a role in establishing these relationships. In crayfish,<br />

chemoreception has been found to play an important role in the<br />

development and maintenance of social hierarchies. During an agonistic<br />

bout, urine is released from the nephropores and is propelled in the<br />

anterior direction toward a conspecific by use of the fan organs<br />

(maxillae and maxillipeds). Conspecifics detect chemical signals with<br />

two pairs of antennules. Previous work has been done on the frequency<br />

and duration of urine releases during agonistic bouts between crayfish.<br />

The results show that dominant crayfish have a longer duration and<br />

more frequent release of urine than subordinates. However, there may<br />

also be differences in the chemical composition of urine in dominant<br />

and subordinate animals, perhaps due to intrinsic variability. This<br />

project used HPLC to examine the different chemical components<br />

present within the urine of dominant and subordinate individuals. The<br />

social status and previous social experience of crayfish altered the<br />

presence or absence of chemical cues within the urine. In addition, the<br />

quantity of urine released differed between dominant and subordinate<br />

crayfish. There<strong>for</strong>e, social status alters the composition of chemicals<br />

within urine. Consequently, it appears as if crayfish can alter the<br />

behavior of conspecifics by releasing urine during a fight, and that this<br />

urine may be a true indicator of social status and fighting ability.<br />

61<br />

236 Poster [ ] Social Odors and Behavior<br />

THE UTILIZATION OF THE MAJOR CHELAE BY MALE<br />

CRAYFISH (ORCONECTES RUSTICUS) FOR DETECTING<br />

FEMALE PHEROMONES<br />

Belanger R.M. 1, Moore P.A. 1 1J.P. Scott Center <strong>for</strong> Neuroscience,<br />

Mind & Behavior, Bowling Green State University, Bowling Green, OH<br />

The abundance and spatial distribution of the sensory hairs of the<br />

major chelae of the crayfish (Orconectes rusticus) are dependent upon<br />

the sex and reproductive <strong>for</strong>m of the crayfish. It has been previously<br />

demonstrated that the major chelae may have chemosensory abilities<br />

that function to detect pheromones or mating signals. Given these<br />

previous results, we determined if <strong>for</strong>m I (reproductive) male crayfish<br />

use their major chelae to detect potential female pheromones in order to<br />

locate potential mates. This study provides a link between the previous<br />

morphological studies and reproductive behavior. We videotaped and<br />

analyzed the behavioral reactions of <strong>for</strong>m I males (N=20) to female<br />

conditioned water (N=6), male conditioned water (N=6), and control<br />

water. Following this, we used males that had their chelae sensory<br />

deprived (lesioned) by coating them with super glue. We found that<br />

unlesioned <strong>for</strong>m I males responded significantly more (P = 0.02) to<br />

female odor than males with lesioned chelae (P = 0.61). These results<br />

clearly show that the major chelae serve a chemosensory function that<br />

is related to distinguishing sex in crayfish. It may be possible that the<br />

chelae also play a chemosensory role in mate localization and courtship<br />

behavior.<br />

237 Poster [ ] Social Odors and Behavior<br />

INDIVIDUAL RECOGNITION IN THE LOBSTER, HOMARUS<br />

AMERICANUS: THE LOSER REMEMBERS<br />

Steinbach M.A. 1, Atema J. 1 1Biology, Boston University, Woods Hole,<br />

MA<br />

Lobsters can distinguish conspecifics individually by odor, and use<br />

this in<strong>for</strong>mation, carried in the urine, to establish their social structure<br />

through a series of agonistic encounters. In the first encounter,<br />

unfamiliar lobsters learn the individual odortype of their opponents. In<br />

second and subsequent encounters, they do not engage in highly<br />

aggressive interactions. In this study, we questioned which of an<br />

agonistic pair makes the decision not to fight a second time. We paired<br />

male lobsters in two successive boxing matches. Be<strong>for</strong>e the second<br />

fight, we disabled the critical antennular chemoreceptors of either the<br />

winner or the loser of fight one. The effects of the lesion on the<br />

behavior of both animals as well as the overall characteristics of the<br />

fight were recorded. Results show that when the subordinate´s<br />

chemoreceptors are disabled and the dominant remains intact, all<br />

behaviors and fight characteristics remain largely the same in fight two<br />

as in fight one. In two cases, the lesioned loser of fight one beat his<br />

dominant in fight two, thus overturning their dominance relationship.<br />

When the winner´s nose is disabled and the loser remains intact,<br />

however, the duration of the fight and all other measures of aggression<br />

decrease significantly <strong>for</strong> both winner and loser. These findings<br />

confirm that the loser of a fight determines the intensity of subsequent<br />

fights, fleeing significantly sooner and more often, thereby eliciting less<br />

aggression from the winner.


238 Poster [ ] Social Odors and Behavior<br />

CHEMICAL SIGNALS AND CHEMOSENSORY PATHWAYS<br />

INVOLVED IN SPINY LOBSTER SHELTERING BEHAVIOR<br />

Horner A.J. 1, Nickles S.P. 1, Derby C.D. 1 1Biology, Georgia State<br />

University, Atlanta, GA<br />

The chemosensory system of the Caribbean spiny lobster (Panulirus<br />

argus) is organized into two parallel pathways that originate in different<br />

populations of antennular sensilla and project to specific neuropils in<br />

the brain. Although unique roles <strong>for</strong> each pathway in olfactory mediated<br />

behaviors have not been described in lobsters, work on other<br />

crustaceans suggests that the pathways may have different roles in the<br />

detection of intraspecific signals. Caribbean spiny lobsters are<br />

gregarious animals that often shelter together in communal dens.<br />

Several previous studies have demonstrated that this aggregation<br />

behavior is mediated by chemical signals released from sheltering<br />

conspecifics. However, the nature of the aggregation signal and the<br />

chemosensory pathways involved in its detection are currently<br />

unknown. We developed a shelter choice assay in a large seawater<br />

flume and examined the sheltering behavior of lobsters in response to a<br />

range of odorants including diluted conspecific urine (pheromone) and<br />

its chemical fractions, shrimp extract (food), and octopus odor<br />

(predator). Lobsters sheltered preferentially with diluted conspecific<br />

urine, showed no preference <strong>for</strong> the shrimp odor and avoided shelters<br />

with octopus odor. Thus sheltering behavior is specific to conspecific<br />

odors, and the aggregation signal is contained within urine. At present<br />

we are attempting to chemically characterize the aggregation<br />

pheromone, and we are also examining the importance of each of the<br />

chemosensory pathways in this behavior through selective ablation of<br />

different populations of antennular sensilla. Supported by NSF IBN-<br />

0077474<br />

239 Poster [ ] Social Odors and Behavior<br />

IN SEARCH OF SEX PHEROMONES IN BLUE CRABS<br />

Kamio M. 1, Kubanek J. 2, Derby C.D. 1 1Biology, Georgia State<br />

University, Atlanta, GA; 2Biology, Georgia Institute of Technology,<br />

Atlanta, GA<br />

Blue crab Callinectes sapidus premolt females release a sex<br />

pheromone in their urine. Males detect this pheromone using antennular<br />

sensors, resulting in mating behaviors that include precopulatory<br />

display (rhythmic waving of swimming legs) and grabbing and<br />

guarding the female. Male crab also releases a sex pheromone that<br />

attracts premolt females (R.A Gleeson in Crustacean Sexual Biology,<br />

Columbia Univ. Press, 1991). The molecular identity of these<br />

pheromones remains unknown. The goal of our study is to identify<br />

these molecules using bioassay guided fractionation and analysis of<br />

differences in the composition of male and female urine. We collected<br />

female and male urine and separated them by ultrafiltration into three<br />

molecular weight fractions: 1000 Da (large). Small and middle fractions of female urine<br />

induced male precopulatory display as well as standing high on legs,<br />

spreading chelae, and grasping. These same size fractions of male urine<br />

induced males to per<strong>for</strong>m all of these behaviors except precopulatory<br />

display. These results indicate that female urine contains a sex-specific<br />

pheromone, and male crab urine contains other chemicals such as<br />

species- or male-specific odors that stimulate agonistic behavior in<br />

other males, or even non-specific odors. Our results also show that the<br />

sex pheromone is either a single molecule ca. 500 Da or mixture of<br />

molecules including some


242 Poster [ ] Social Odors and Behavior<br />

NEW INSIGHTS ON THE SOCIAL STRUCTURE AND ODOR<br />

FUNCTION OF A TANGERINE-SCENTED SEABIRD<br />

Kett L. 1, Hagelin J. 1, Rasmussen L. 2 1Department of Biology,<br />

Swarthmore College, Swarthmore, PA; 2Biochemistry and Molecular<br />

Biology, Oregon Graduate Institute, Beaverton, OR<br />

The crested auklet (Aethia cristatella ) is an arctic seabird that emits<br />

a tangerine-scented odor during breeding (Hagelin et al., 2003). Birds<br />

rub their beaks in the scented nape of a display partner, providing a<br />

possible mechanism <strong>for</strong> odor assessment. The exact social function of<br />

crested auklet odor is, as yet, unknown. Two possible hypotheses<br />

include: (1) odor correlates with social rank, such as dominance status,<br />

or (2) odor acts as a sexual ornament that attracts members of the<br />

opposite sex. We analyzed patterns of aggression and courtship within a<br />

captive population of 14 birds. We also collected chemical samples<br />

from both the oil gland and freshly clipped feathers. Behavioral analysis<br />

revealed a strong linear hierarchy within our population. Among<br />

females, social (dominance) rank was positively related to the degree to<br />

which a female associated with a mate. Birds also mated assortatively<br />

with respect to body size. Chemical analyses revealed that both oil<br />

glands and feathers contained compounds characteristic of the auklet´s<br />

seasonal scent (e.g. cis-4-decenal and octanal; Hagelin et al., 2003).<br />

From the samples of top ranking females we identified more than 25<br />

volatile compounds in oil gland secretions, including a variety of C -C 8 10<br />

acids, aldehydes and alcohols. An analysis of the relative concentration<br />

of specific compounds versus social rank is currently underway. Such a<br />

comparison promises to reveal patterns related to odor function, given<br />

the surprisingly linear social hierarchy of our captive population.<br />

Funding <strong>for</strong> our research was provided, in part, through the generous<br />

support of the Aquarium of the Pacific, Long Beach, Cali<strong>for</strong>nia.<br />

243 Poster [ ] Social Odors and Behavior<br />

BEHAVIORAL AND PHYSIOLOGICAL RESPONSES TO A<br />

PUTATIVE ALARM ODOR IN EUROPEAN STARLINGS<br />

(STURNUS VULGARIS).<br />

Leininger E.C. 1, Hile A. 2, Hagelin J. 1 1Biology, Swarthmore College,<br />

Swarthmore, PA; 2USDA/APHIS/WS/NWRC/Monell Chemical Senses<br />

Center, Phiadelphia, PA<br />

Wild European Starlings (Sturnus vulgaris) produce a pungent odor<br />

that is detectable to humans when birds are held in captivity. Despite<br />

growing evidence <strong>for</strong> chemosignals in birds, alarm odors have not been<br />

investigated. We tested whether fecal odor of stressed European<br />

Starlings produced a behavioral or physiological alarm response in nonstressed<br />

individuals. We videotaped focal birds (n=18) while exposed<br />

to one of three chemical stimuli: (1) fecal odor from a stressed<br />

conspecific, (2) fecal odor from an unstressed conspecific, and (3) plain<br />

air. Fecal samples of focal birds were also collected be<strong>for</strong>e and after<br />

exposure to each treatment, and assayed <strong>for</strong> a common stress hormone<br />

(corticosterone [CORT]). Analyses of gaping and breathing rate,<br />

common distress behaviors in birds, are currently underway. We<br />

detected no significant difference between the three treatments in the<br />

frequency of other behaviors such as hopping, (p=0.58) preening<br />

(p=0.93), or feather fluffing (p=0.51). With regard to physiology,<br />

starlings tended to exhibit less of a net increase in fecal CORT levels<br />

after exposure to stressed conspecific odor, compared to other chemical<br />

treatments (p=0.097). Such a result suggests that starlings might<br />

experience a physiological change after exposure to the scent of<br />

stressed conspecifics. Though our preliminary study failed to meet<br />

statistical significance, it is an intriguing first step that highlights the<br />

need <strong>for</strong> additional, careful study of avian chemical signals.<br />

Funding: HHMI Summer Research Stipend from Swarthmore<br />

College (EL); USDA/ APHIS/ WS, NWRC, and Monell Chemical<br />

Senses Center (AGH).<br />

63<br />

244 Poster [ ] Social Odors and Behavior<br />

THE INFLUENCE OF CONTEXT ON FEMALE MHC-BASED<br />

MATE CHOICE<br />

Shaw - Taylor E.E. 1, Mcclintock M. 2 1Pyschology, University of<br />

Chicago, Chicago, IL; 2Institute <strong>for</strong> Mind and Biology, University of<br />

Chicago, Chicago, IL<br />

The primary function of the molecules of the major<br />

histocompatibility complex, the MHC, is to enable antigen presentation<br />

to T-cells, thereby initiating an immune response. The MHC also<br />

influences mating behavior such that individuals choose mates who<br />

express MHC alleles that are different from their own. MHC-based<br />

disassortative mate choice has been shown in mice, fish, and humans.<br />

Females more so than males may mate disassortatively with respect to<br />

the MHC because of potential immune and fitness benefits that this<br />

genetic diversity provides offspring. However, reports from some<br />

MHC-based odor studies, wherein females were presented with the<br />

odors from males possessing many different MHC alleles, did not<br />

indicate preferences <strong>for</strong> the most MHC differences. Rather, females<br />

preferred the odors of males with whom they shared a few of the same<br />

alleles, suggesting that the preferred MHC difference is relative to the<br />

female´s own MHC diversity. We previously tested this hypothesis and<br />

presented data consistent with preference <strong>for</strong> differences—the greater<br />

the MHC difference, the greater the preference. However, we also<br />

showed that the direction of preference depended on the type of choice<br />

being made—whether females were choosing between a male identical<br />

in MHC and one different or between two males that were both<br />

different in MHC, in number of alleles. To further investigate the<br />

influence of context, we have expanded the study with more subjects<br />

and another MHC haplotype. We present data that female rats (Rattus<br />

norvegicus) consider the context of choice, as indicated by male MHC<br />

differences, when making mating decisions.<br />

Supported by an NIMH MERIT Award to MKM and Hinds Research<br />

Fund.<br />

245 Poster [ ] Social Odors and Behavior<br />

THE SCENT OF FRIENDSHIP: HIGH SCHOOL STUDENTS<br />

RESEARCH THE MYSTERIES OF HUMAN ODOR<br />

RECOGNITION<br />

Olsson S.B. 1, Barnard J. 2, Turri L. 2 1Neurobiology and Behavior,<br />

Cornell University, Geneva, NY; 2Biology, Geneva High School,<br />

Geneva, NY<br />

Though several studies have examined the effect of human odor on<br />

kin recognition and mate choice, few have focused on the impact of<br />

familiarity on recognition of non-relatives. As part of a program<br />

designed to engage students in scientific research, 55 10th grade and<br />

Advanced Placement biology students researched, planned, and<br />

implemented a project to analyze the effect of odor on human<br />

recognition of, and preference <strong>for</strong>, close friends, gender, and self. Each<br />

student and friend of their choosing wore a T-shirt <strong>for</strong> three consecutive<br />

nights. During that time, subjects were controlled <strong>for</strong> exposure to<br />

extraneous perfumes, household odors, and other humans. The students<br />

were then asked to smell a series of shirts and evaluate them with<br />

respect to pleasantness. Students were also asked to identify the two<br />

shirts belonging to themselves and their friend, and determine the<br />

gender of each shirt. Results of this testing will be presented along with<br />

a discussion of its implications <strong>for</strong> human social behavior. This research<br />

is supported by the NSF Graduate Teaching Fellows in K-12 Education<br />

Program and Cornell University through the Cornell Science Inquiry<br />

Partnerships Program.


246 Slide [ ] Olfactory Behavior & Psychophysics<br />

FORAGING IN A COMPLEX CHEMICAL LANDSCAPE: DOM<br />

FROM ELEVATED CO2 DETRITUS AND ITS IMPACT ON<br />

CRAYFISH ORIENTATION TO A FOOD SOURCE<br />

Adams J. 1, Moore P.A. 1 1Biological <strong>Sciences</strong>, Bowling Green State<br />

University, Bowling Green, OH<br />

Crayfish must locate food resources in a chemically complex<br />

environment where chemicals from various sources interact and mix<br />

with one another. We tested whether an additional chemical food source<br />

of leachate from detritus (dissolved organic matter, DOM) produced at<br />

ambient (AMB) or elevated (ELEV) atmospheric CO and presented at<br />

2<br />

two different concentrations would affect crayfish orientation behavior.<br />

Crayfish (Orconectes virilis) orientation was observed in a recirculating<br />

flume where a fish odor source was placed downstream of a DOM odor<br />

source. Preliminary analysis of the chemical signal demonstrated that<br />

the fine-scale spatiotemporal structure of the odor plume was different<br />

upon introduction of a background DOM odor. Behavioral treatments<br />

were as follows: 1) CON (Control; no DOM present), 2) AMB-Low<br />

(AMB DOM present at 3 mg/L), 3) AMB-High (AMB DOM present at<br />

6 mg/L), 4) ELEV-Low (ELEV DOM present at 3 mg/L), and 5)<br />

ELEV-High (ELEV DOM present at 6 mg/L). Crayfish in the AMB-<br />

High treatment were more successful in locating the source. In contrast,<br />

animals in the ELEV-High treatment had higher turning angles and<br />

heading angles than all other treatments. No differences were found <strong>for</strong><br />

temporal parameters of orientation. Overall, these results indicate that<br />

crayfish orientation to a fish odor source is affected by the presence of<br />

DOM from detritus when it is present at the high end of a natural range<br />

of DOM concentration, perhaps resulting in a chemical “cocktail party”<br />

effect. This research was funded by the NSF/IGERT BART program<br />

(JAA) and NSF DAB 9874608 (PAM).<br />

247 Slide [ ] Olfactory Behavior & Psychophysics<br />

CHEMICALLY-INDUCED ANTENNULAR GROOMING IN<br />

THE SPINY LOBSTER, PANULIRUS ARGUS, IS MEDIATED<br />

BY NON-OLFACTORY SENSILLA<br />

Schmidt M. 1, Derby C.D. 1 1Biology, Georgia State University, Atlanta,<br />

GA<br />

In decapod crustaceans, the 1st antennae bearing olfactory and other<br />

chemo-mechanosensory sensilla are groomed by mouthpart appendages<br />

in a stereotyped behavioral pattern. In the spiny lobster, Panulirus<br />

argus, antennular grooming behavior (AGB) is elicited by L-glutamate,<br />

and ablation experiments led to the conclusion that AGB is mediated by<br />

olfactory sensilla (aesthetascs) located on the lateral flagella of the 1st<br />

antennae (Wroblewska et al., Chem. Senses 27:769-778, 2002). The<br />

aim of our study was to examine this conclusion since the aesthetascs<br />

are closely associated with other sensilla, the asymmetric setae, which<br />

had also been eliminated. In two independent sets of experiments, we<br />

first showed that intact animals responded with AGB upon stimulation<br />

with 0.5 mM L-glutamate, then removed either all asymmetric setae (N<br />

= 8 / N = 7) or all aesthetascs (N = 8 / N = 6). Aesthetasc ablation did<br />

not reduce AGB, whereas ablation of asymmetric setae almost<br />

completely eliminated it. A 3rd set of experiments showed that<br />

aesthetasc ablation also has no positive effect on AGB. Electron and<br />

confocal microscopy suggest that asymmetric setae are bimodal chemomechanosensory<br />

sensilla, defined by the presence of a terminal pore<br />

and a scolopale. We conclude that AGB is elicited by chemosensory<br />

input provided by asymmetric setae and that it is not mediated by the<br />

olfactory pathway but by a parallel chemo- and mechanosensory<br />

pathway constituted by the lateral antennular neuropils (Schmidt &<br />

Ache, J Comp Physiol A 178:579-604, 1996). Supported by NSF (IBN-<br />

0077474)<br />

64<br />

248 Slide [ ] Olfactory Behavior & Psychophysics<br />

DISCRIMINATION BETWEEN ENANTIOMERS OF CARVONE<br />

AND TERPINEN-4-OL ODORANTS IN NORMAL RATS AND<br />

THOSE WITH LESIONS OF THE OLFACTORY BULBS<br />

Slotnick B. 1, Mcbride K. 2 1Psychology, University of South Florida,<br />

Tampa, FL; 2Psychology, American University, washington, DC<br />

Rats trained on an operant conditioning task were used to assess<br />

whether the enantiomers of terpinen-4-ol, odorants that activate very<br />

similar areas of the olfactory bulb are more difficult to discriminate<br />

than those of carvone, odorants that activate different areas of the<br />

olfactory bulb, and to determine the effects of disrupting identified<br />

bulbar sites activated by these odorants. In psychophysical tests in<br />

which odor concentration was gradually reduced to near threshold<br />

levels, normal rats discriminated between the enantiomers of terpinen-<br />

4-ol and of carvone equally well. Olfactory bulb lesions that removed<br />

the majority of bulbar glomeruli activated by these odorants (as<br />

demonstrated in prior olfactory bulb studies using optical imaging and<br />

2-deoxyglucose) resulted in increased detection thresholds but little or<br />

no deficits in discriminating between supra threshold concentrations of<br />

the enantiomers. These results fail to confirm predictions based on<br />

maps of bulbar activity that enantiomers of terpinen-4-ol should be<br />

more difficult to discriminate than those of carvone and that the ability<br />

to discriminate between enantiomers of an odorant are based on<br />

differences in patterns of bulbar activation produced by the odorants.<br />

249 Slide [ ] Olfactory Behavior & Psychophysics<br />

SIZE AND NUMBERS DON´T MATTER (THAT MUCH) -<br />

RELATIVE SIZE OF OLFACTORY BRAIN STRUCTURES AND<br />

NUMBER OF FUNCTIONAL OLFACTORY RECEPTOR<br />

GENES ARE POOR PREDICTORS OF OLFACTORY<br />

PERFORMANCE<br />

Laska M. 1 1University of Munich, Munich, Germany<br />

Primates are typically regarded as „microsmatic“ animals. This view,<br />

however, is only based on an interpretation of neuroanatomical and<br />

recent genetic findings and not on physiological evidence. We<br />

determined olfactory detection thresholds in squirrel monkeys and<br />

pigtail macaques <strong>for</strong> more than 40 odorants from different chemical<br />

classes and compared their per<strong>for</strong>mance to that of human subjects and<br />

nonprimate mammals. We found that – contrary to the traditional view -<br />

human subjects do not generally per<strong>for</strong>m poorer than monkeys, and Old<br />

World primates do not generally show higher threshold values than<br />

New World primates. Furthermore, human and nonhuman primates<br />

show an olfactory sensitivity which <strong>for</strong> several substances matches or<br />

even is markedly better than that of species believed to be<br />

„macrosmatic“ such as dogs, rats, or mice. Similarly, human and<br />

nonhuman primates do not generally per<strong>for</strong>m poorer in discriminating<br />

between structurally related odorants compared to nonprimate<br />

mammals and insects.We conclude that between-species comparisons<br />

of neuroanatomical features or of the number of functional olfactory<br />

receptor genes are poor predictors of olfactory per<strong>for</strong>mance.<br />

Differences in olfactory sensitivity and discrimination abilities within<br />

and between species seem to reflect evolutionary adaptations to<br />

ecological niches.


250 Slide [ ] Olfactory Behavior & Psychophysics<br />

A PSYCHOPHYSICAL TEST OF THE VIBRATION THEORY<br />

OF OLFACTION<br />

Keller A. 1, Vosshall L.B. 1 1Laboratory of Neurogenetics and Behavior,<br />

Rockefeller University, New York, NY<br />

At present no satisfactory theory exists to explain why a given<br />

molecule has a particular smell. A recent book about the physiologist<br />

Luca Turin has generated new interest in the theory that the smell of a<br />

molecule is determined by its intramolecular vibrations rather than by<br />

its shape. We present the first psychophysical experiments in humans<br />

that test key predictions of this theory. The results suggest that<br />

molecular vibrations alone cannot explain the perceived smell of a<br />

chemical. Specifically, we have found that: (i) in a component<br />

identification task no vanilla odor character was detected in the mixture<br />

of benzaldehyde and guaiacol (ii) odor similarity ratings did not reveal<br />

that even and odd numbered aldehydes <strong>for</strong>m two odor classes and (iii)<br />

naive subjects who could easily discriminate the smell of two molecules<br />

that differ in shape but not in molecular vibration failed to discriminate<br />

two molecules with similar shape but different molecular vibrations in<br />

three different experimental paradigms (similarity rating, duo-trio test,<br />

triangle test). Taken together our findings are consistent with the idea<br />

that the smell of a molecule is determined by its shape but we found no<br />

evidence that the smell of a molecule is influenced by its vibrational<br />

properties.<br />

251 Slide [ ] Olfactory Behavior & Psychophysics<br />

FUNCTIONAL CONNECTIVITY OF THE HIPPOCAMPUS<br />

DURING AN OLFACTORY TASK: DIFFERENCES OBSERVED<br />

BETWEEN YOUNG AND ELDERLY<br />

Calhoun-Haney R. 1, Ferdon S. 2, Barbara C. 2, Murphy C. 2 1Clinical<br />

Psychology, San Diego State University, San Diego, CA; 2Psychology,<br />

San Diego State University, San Diego, CA<br />

Interest in the role of the hippocampus (HIP) <strong>for</strong> olfactory function<br />

has increased since the discovery that initial neuropathological changes<br />

indicative of Alzheimer's Disease (AD) occur in mesial temporal<br />

olfactory regions. The present study investigated differences in the<br />

functional relationship between the HIP and classic olfactory regions in<br />

young and elderly adults in response to odor stimulation. Activation in<br />

reference voxels located in left and right hippocampi were separately<br />

averaged and correlation coefficients between these and all other brain<br />

voxels were calculated. Group analyses were also per<strong>for</strong>med in order to<br />

detect brain areas with significant differences in their strength of<br />

functional association to the hippocampi. For the young, both left and<br />

right hippocampi demonstrated strong functional associations with<br />

classic olfactory areas while the elderly exhibited a reduced number of<br />

functional associations. In addition, while both right and left<br />

hippocampi demonstrated functional connectivity with ipsilateral<br />

orbital frontal cortices in the young, functional connectivity was not<br />

observed between the left HIP and left orbital frontal cortex in the<br />

elderly. However, a notably larger area of the right orbital frontal cortex<br />

displayed connectivity with the right HIP, suggesting the possibility of<br />

a compensatory process in the elderly. Group analyses further<br />

demonstrated age differences in strength of functional associations<br />

among regions. The present study illustrates the effect of aging on<br />

hippocampal function and provides a foundation <strong>for</strong> further<br />

understanding of its dysfunction when affected by a neurodegenerative<br />

disease such as AD. Supported by NIH grant AG04085 to CM.<br />

65<br />

252 Slide [ ] Olfactory Behavior & Psychophysics<br />

IMPACT OF THE CHEMICAL SENSES ON AUGMENTING<br />

MEMORY, ATTENTION, REACTION TIME, PROBLEM<br />

SOLVING, AND RESPONSE VARIABILITY: THE<br />

DIFFERENTIAL ROLE OF RETRONASAL VERSUS<br />

ORTHONASAL ODORANT ADMINISTRATION<br />

Zoladz P. 1, Raudenbush B. 1, Lilley S. 1 1Psychology, Wheeling Jesuit<br />

University, Wheeling, WV<br />

Past research has consistently noted a significant interplay between<br />

odors and human behavior. Multiple studies have shown that the<br />

administration of particular odorants can enhance athletic per<strong>for</strong>mance,<br />

sleep, pain tolerance, mood, and cognitive processing. In addition,<br />

odorants have a differential effect on human behavior, dependent upon<br />

route of administration (retronasal vs. orthonasal). The present study<br />

examined the differential effects of odorants administered retronasally<br />

and orthonasally on cognitive per<strong>for</strong>mance. During Phase I, 31<br />

participants completed cognitive tasks on a computer-based program<br />

(Impact®) under five "chewing gum" conditions (no gum, flavorless<br />

gum, peppermint gum, cinnamon gum, and cherry gum). During Phase<br />

II, 39 participants completed the same cognitive tasks under four<br />

odorant conditions (no odor, peppermint odor, jasmine odor, and<br />

cinnamon odor). Participants also completed pre- and post-test<br />

assessments of mood, and rated their perception of the required<br />

workload. Results revealed a task-dependent relationship between odors<br />

and the enhancement of cognitive processing. Cinnamon, administered<br />

retronasally or orthonasally, improved participants' scores on tasks<br />

related to attentional processes, virtual recognition memory, working<br />

memory, and visual-motor response speed. Implications of the present<br />

study are most promising in providing a non-pharmacological adjunct<br />

to enhancing cognition in the elderly, individuals with test-anxiety, and<br />

perhaps even patients with diseases that lead to cognitive decline. This<br />

study was funded by a grant from Psi Chi to P. Zoladz.<br />

253 Slide [ ] Olfactory Behavior & Psychophysics<br />

THE MAGIC NUMBER 3 APPLIES TO COMPONENTS<br />

IDENTIFIED IN COMPLEX ODOR-TASTE MIXTURES<br />

Laing D. 1, Marshall K. 1, Jinks A. 1, Hutchinson I. 1 1Centre For<br />

Advanced Food Research, University of Western Sydney, Sydney,<br />

Australia<br />

Humans can identify up to 3 components in taste or odour mixtures<br />

(Laing & Francis, 1989; Laing et al., 2002), but their capacity to<br />

analyse multi-component odour-taste mixtures is unknown. With binary<br />

mixtures both components are usually perceived, however, only tastants<br />

were identified in 3-component mixtures containing one odorant.<br />

(Laing et al., 2002), suggesting taste may dominate smell in odor-taste<br />

mixtures. Here, the aim was to determine the number of identifiable<br />

components in complex odor-taste mixtures and investigate the report<br />

of dominance of taste over smell. 43 subjects were trained to identify<br />

'equi' intense aqueous solutions of the tastants sucrose, sodium choride<br />

and citric acid, and the odorants cinnamaldehyde (cinnamon), cis 3hexenol<br />

(grass-like) and 2-pentanone (like nail polish remover). Over 2<br />

test sessions they were asked to identify the components of 36 mixtures<br />

containing from 1 to 6 components presented in random order. Stimuli<br />

were sampled by mouth using a straw sited in a hole in the lid of a 30<br />

ml plastic cup. Subjects readily identified components in 1 and 2<br />

component stimuli, but with 3-6 component mixtures the mean number<br />

identified was 3. Clearly, increasing the number of modalities in a<br />

mixture did not increase the number of components identified. As<br />

reported earlier, tastes were more readily identified than odors in<br />

mixtures. The limit of 3 components strongly suggests the limiting<br />

factor is working memory, the memory that is used to rapidly identify a<br />

stimulus and initiate a response within a second or two.


254 Symposium [ ] Non-neuronal Cells of the Olfactory<br />

System in Development<br />

SUSTENTACULAR CELLS - MORE ACTIVE THAN WE EVER<br />

IMAGINED<br />

Hegg C. 1, Vogalis F. 1, Lucero M. 1 1Physiology, University of Utah, Salt<br />

Lake City, UT<br />

Sustentacular cells are morphologically polarized and have structural<br />

features that allude to functions of secretion, absorption, phagocytosis,<br />

maintenance of extracellular ionic gradients, metabolism of noxious<br />

chemicals, and regulation of cell turnover. Here, we review the well<br />

characterized morphology and ultrastructure of mammalian<br />

sustentacular cells and present data investigating their dynamic activity.<br />

We show, using a mouse olfactory epithelium slice model, that<br />

sustentacular cells have voltage-gated Na + and K + channels and are<br />

capable of evoking rapid, robust increases in intracellular calcium in<br />

response to P2Y purinergic receptor activation. In addition,<br />

sustentacular cells propagate waves of calcium oscillations initiated by<br />

either endogenous release or exogenous application of ATP or other<br />

putative neurotransmitters. Oscillations in intracellular calcium may<br />

govern secretion, proliferation and development in sustentacular cells,<br />

and, via calcium dependent exocytosis, may provide chemical signals to<br />

basal cells, neuronal precursors, or neurons. Recent discoveries of<br />

voltage-gated channels, receptors, and transmitter release in both<br />

peripheral and central glial cells suggest direct communication between<br />

neurons and glia. The identification of similar components in<br />

sustentacular cells suggests that, like their glial counterparts, they are<br />

capable of rapid communication between themselves and the neural<br />

elements of the olfactory epithelium. Funded by NIDCD DC02994<br />

(MTL) and DC04953 (CCH).<br />

255 Symposium [ ] Non-neuronal Cells of the Olfactory<br />

System in Development<br />

A GLIA - AXON PAS DE DEUX UNDERLIES OLFACTORY<br />

RECEPTOR AXON SORTING.<br />

Oland L.A. 1 1A.R.L. Division of Neurobiology, University of Arizona,<br />

Tucson, AZ<br />

The problem: Olfactory receptor neurons expressing the same<br />

olfactory specificity are not topographically ordered within the receptor<br />

epithelium, but their axons extend to just one or a pair of glomeruli in<br />

the olfactory bulb. Axon sorting is thus a critical process that must<br />

occur somewhere along the olfactory pathway. In the moth Manduca<br />

sexta, sorting takes place not in the nerve layer that circumscribes the<br />

olfactory neuropil, but rather in a discrete region of the nerve that we<br />

have called the axon sorting zone. This sorting zone has three<br />

important features: it is densely populated by glial cells, the glial cells<br />

can be killed without killing neurons, and it can easily be surgically<br />

isolated <strong>for</strong> various in vivo and in vitro experimental manipulations.<br />

Recent experiments have provided clear evidence <strong>for</strong> reciprocal<br />

interactions between the ingrowing receptor axons and the glial cells<br />

that affect glial cell proliferation, axon fasciculation, and growth cone<br />

behavior, and also are beginning to reveal the underlying molecular<br />

players.<br />

66<br />

256 Symposium [ ] Non-neuronal Cells of the Olfactory<br />

System in Development<br />

SORTING AND GLIAL-NEURONAL INTERACTIONS IN THE<br />

OLFACTORY NERVE LAYER<br />

Treloar H.B. 1, Akins M. 1, Iwema C. 1, Dodds T. 1, Greer C.A. 2<br />

1Neurosurgery, Yale University, New Haven, CT; 2Neurobiology, Yale<br />

University, New Haven, CT<br />

Olfactory Sensory neurons (OSNs), are generated throughout life and<br />

extend axons that <strong>for</strong>m fascicles within the olfactory nerve (ON). As<br />

axons enter the olfactory nerve layer (ONL) of the olfactory bulb (OB),<br />

they reorganize be<strong>for</strong>e synapsing in glomeruli. Each OSN expresses<br />

one odorant receptor (OR) from a family of approx. 1000 ORs. Neurons<br />

expressing the same OR target specific glomeruli. Thus, OSN axons<br />

face unique challenges when sorting. First, axons from neurons located<br />

widely within the OE must identify common targets, taking varied<br />

trajectories and presumably using multiple guidance cues. Second, as<br />

neurons are continually regenerated throughout life, axons must<br />

navigate a complex meshwork of pre-existing axons to identify target<br />

glomeruli. Throughout this process the unymyelinated OSN axons<br />

maintain intimate contact with a unique population of glia, the olfactory<br />

ensheathing cells (OECs). The morphological and molecular properties<br />

of OECs vary along the length of the axonal projection, suggesting that<br />

subsets of OECs subserve different functions. In the ON and peripheral<br />

regions of the ONL, OECs tightly bundle OSN axons and express p75.<br />

However, as axons approach glomeruli and radically change their<br />

trajectories, the population of glia they interact with changes. The<br />

OECs found in this region express NPY but not p75. OSN axons also<br />

encounter astrocytic processes within this region which emanate<br />

radially into the ONL from glomeruli. It is likely that a spatiotemporal<br />

combination of cell surface axon-axon and axon-glia interactions<br />

underlie glomerular targeting in the OB.<br />

NIH DC005706, DC00210


257 Symposium [ ] Non-neuronal Cells of the Olfactory<br />

System in Development<br />

LOSING THE PATH; CELL MIGRATION IN A CHANGING<br />

FOREBRAIN<br />

Demarchis S. 1, Rossi F. 2, Fasolo A. 1, Puche A.C. 3 1Dept. Human &<br />

Animal Biol., Univ. of Turino, Turino, ., Italy; 2Dept. Neuroscience,<br />

Univ. of Turino, Turino, ., Italy; 3Dept. Anatomy and Neurobiol., Univ.<br />

of Maryland, Baltimore, MD<br />

The presence of a germinal layer and the capacity to generate<br />

neurons, once thought restricted to the embryonic brain, persists in the<br />

<strong>for</strong>ebrain of postnatal and adult mammals. The olfactory bulb (OB),<br />

unlike the cortex and other regions, continues to receive new neurons<br />

throughout life. Despite dramatic changes in the structure and cellular<br />

organization of the <strong>for</strong>ebrain from embryo to adult, these newly<br />

generated cells still migrate from germinal regions to specific targets. In<br />

early embryos, the OB has a cortical-like organization with radial glia<br />

spanning the ventricle to pia. Radial glia are likely to provide a<br />

migratory scaffold <strong>for</strong> cells exiting the ventricular germinal layer and<br />

migrating radially. As development progresses numerous cells are<br />

generated from a stem cell pool located in the lateral ventricle<br />

subventricular zone (SVZ). SVZ-derived progenitor cells migrate into<br />

the OB following the rostral migratory stream (RMS) and into other<br />

postnatal <strong>for</strong>ebrain structures along different migratory pathways. After<br />

birth radial glia disappear and reorganize into a series of glial tubes<br />

along the RMS. However, once radial glia in the bulb disappear there is<br />

no longer a clear radial scaffold <strong>for</strong> cells to follow when migrating<br />

radially. Such modifications of the non-neuronal migratory<br />

environment are paralleled by changes in lineage characteristics of SVZ<br />

progenitors. Namely, progenitors that leave the SVZ at different<br />

developmental times generate different types of mature OB neurons.<br />

Thus, cell migration in the bulb is a developmentally dynamic process<br />

between immature cells and different non-neuronal cellular elements at<br />

different times. Support: NIDCD DC05739, Compagnia di San Paolo.<br />

and Italian Institute <strong>for</strong> Health Project CS107.1<br />

258 Poster [ ] Chemical Ecology<br />

ORIENTATION TO TEMPORALLY AND SPATIALLY<br />

COMPLEX ODOR SIGNALS IN THE CRAYFISH,<br />

ORCONECTES RUSTICUS<br />

Zulandt T.J. 1, Quinn E. 2, Wolf M. 3, Moore P.A. 3 1Laboratory <strong>for</strong><br />

Sensory Ecology, Bowling Green State University, Bowling Green<br />

Ohio, OH; 2Electrical Engineering, University of Cincinnati,<br />

Cincinnati, OH; 3Biological <strong>Sciences</strong>, Bowling Green State University,<br />

Bowling Green, OH<br />

Organisms in natural environments encounter complex in<strong>for</strong>mation<br />

and must be able to decipher this in<strong>for</strong>mation in order to respond<br />

behaviorally. This complex in<strong>for</strong>mation can be a result of various<br />

aspects of the natural environment such as hydrodynamics, changing<br />

habitats, and conflicting or multiple odor cues. In many habitats,<br />

multiple odor plumes in various spatial configurations often mix and<br />

animals must extract relevant spatial in<strong>for</strong>mation in order to make<br />

appropriate orientation decisions. In this study, we show how the spatial<br />

configuration of multiple odor plumes influences the orientation<br />

abilities of crayfish. By altering the spatial configuration of odor<br />

sources without altering concentration or hydrodynamics it is possible<br />

to understand the underlying mechanisms of orientation. In this study,<br />

we have quantified the hydrodynamics, chemical dynamics and animal<br />

behavior within the artificial stream. Our results show that the different<br />

spatial configuration influences the distribution of odors within our<br />

stream and that these odors influence the subsequent orientation<br />

behavior. Our work indicates that it is the fine-scale distribution of<br />

odors that is regulating the orientation behavior of the crayfish.<br />

67<br />

259 Poster [ ] Chemical Ecology<br />

FLUID DYNAMICS AND CHEMICAL SIGNALS IN THE<br />

CRAYFISH WALKING LEGS<br />

Cook M. 1, Moore P.A. 1 1Biological <strong>Sciences</strong>, Bowling Green State<br />

University, Bowling Green, OH<br />

In aquatic environments, chemical cues are important in many<br />

species <strong>for</strong> finding resources and locating mates. Two processes<br />

disperse these cues: fluid flow and molecular diffusion. Fluid flow is<br />

dominant in spatial scales larger than 10 mm, while molecular diffusion<br />

is dominant at smaller spatial scales. For animals with chemoreceptive<br />

capabilities, sampling the environment is critical to obtain in<strong>for</strong>mation.<br />

Receptor structure morphology is responsible <strong>for</strong> changing the flow<br />

surrounding receptor cells by creating a boundary layer. Fluid that is in<br />

direct contact with the receptor surface does not flow, which is known<br />

as the no-slip condition. The boundary layer acts as a filter, changing<br />

the temporal and spatial structure of the chemical signal arriving to<br />

receptor cells. In this study, we will employ the technique used by<br />

Moore and Atema 1991 to study the fluid dynamics in the<br />

chemosensory chelae and walking legs of the crayfish, Orconectes<br />

rusticus. Our goal is to determine the function of the boundary layer<br />

be<strong>for</strong>e and after chemical in<strong>for</strong>mation has been filtered (i.e. during<br />

flicking) to the receptor cells. The experiment will be conducted in a<br />

recirculating flow tank in which the stimulus will be delivered with a<br />

plastic pipette. High-speed electrochemical recordings will be made<br />

with BAS epsilon system. It is clear that the morphology of these<br />

appendages influence the structure of chemical signals arriving at<br />

receptor cells. These results provide insight into how crayfish perceive<br />

chemical signals and extract in<strong>for</strong>mation from turbulent odor plumes.<br />

260 Poster [ ] Chemical Ecology<br />

THE ROLE THAT BOUNDARY LAYERS AROUND CRAYFISH<br />

SENSORY APPENDAGES ACT AS TEMPORAL FILTERS FOR<br />

ODOR PLUMES<br />

Urban L. 1, Moore P.A. 1 1Biological <strong>Sciences</strong>, Bowling Green State<br />

University, Bowling Green, OH<br />

Sensory systems have a complex task of extracting relevant<br />

in<strong>for</strong>mation from the environment.The interaction between fluid flow<br />

and the morphology of sensory appendages acts as a physical filter and<br />

plays an important role in extracting in<strong>for</strong>mation.This interaction<br />

determines the boundary layer structure which alters the spatial and<br />

temporal distribution of chemical signals arriving at the receptors. This<br />

study was designed to exmaine the microscale fluid and chemical<br />

dynamics around the sensory appendages of the crayfish, Orconectes<br />

rusticus. Previous research has determined the biomechanics of<br />

sampling behaviors of the lateral antennule of the crayfish. In this<br />

study, we have used these biomechanical results with an<br />

electrochemical technique to quantify the role that boundary layers play<br />

around various sensory appendages in filtering the temporal dynamics<br />

of odor plumes.We measured the boundary layer structure and chemical<br />

dynamics of sensory appendages that were mounted in a flow tank<br />

under known flow conditions. We utilized the BAS epsilon system with<br />

electrochemical microelectrodes to record and quantify chemical<br />

dynamics with multiple electrodes around the appendage. The<br />

preliminary findings indicate the majority of the chemical signal flows<br />

around the appendage rather then through it. Boundary layer structure<br />

greatly affects the spatial and temporal nature of chemical signals by<br />

acting as a smoothing filter <strong>for</strong> incoming signals.The exact nature of<br />

this physical filter depends upon the ambient flow speed and the angle<br />

of the sensory appendage relative to flow.This study illustrates the<br />

importance of <strong>for</strong>m and function in chemoreception and in an<br />

organism's ability to detect environmental cues.


261 Slide [ ] Chemical Ecology<br />

FROM ODOR PLUME TO ANTENNULE: DO CRAYFISH<br />

ANTENNULES VARY WITH FLOW HABITAT AS PREDICTED<br />

TO MAXIMIZE ODOR MOLECULE CAPTURE?<br />

Mead K.S. 1 1Biology, Denison University, Granville, OH<br />

Many aquatic crustaceans use water-borne chemical cues in<br />

ecologically critical activities such as finding food, mates, habitat, and<br />

avoiding predators. These chemical cues are present as odor plumes<br />

consisting of fine filaments of highly concentrated odor molecules<br />

interspersed with the surrounding fluid. The structure of the odor<br />

plume, and thus how the plume is encountered by navigating animals, is<br />

affected by factors such as the size-scale of the bottom substrate and<br />

flow conditions such as the mean velocity, turbulence level, and the<br />

gradient of flow speed above the substratum. Several species of Ohio<br />

crayfish (Cambarus spp. and Orconectes spp.) were collected from a<br />

variety of flow habitats including still ponds and turbulent streams.<br />

Since odor plume structure varies according to flow habitat, crayfish<br />

antennules from species living in different flow environments should<br />

have different patterns of aesthetasc arrangements on their filaments, to<br />

best encounter odors in that habitat. I used Image J to measure<br />

structural parameters such as aesthetasc number, length and diameter,<br />

and the ratio of the gap between aesthetasc rows to the aesthetasc<br />

diameter (an indication of odor penetration into the sensor array) from<br />

SEMs on 3 individuals per species. As predicted from odor plume<br />

structure, species from high flow habitats have longer aesthetascs<br />

(p=0.009; rsPc=0.684) and a smaller ratio of the gap between aesthetasc<br />

bundles to the aesthetasc diameter (p=0.092; rsPc=0.908) than species<br />

from low flow habitats). Funding: Denison University Research<br />

Foundation.<br />

262 Poster [ ] Chemical Ecology<br />

DO MOVEMENTS OF HONEYBEE ANTENNAE ENHANCE<br />

CAPTURE OF ODORANTS?<br />

Miller G. 1, Loudon C. 1, Smith B.H. 2 1Entomology, University of<br />

Kansas, Lawrence, KS; 2Entomology, Ohio State University, Columbus,<br />

OH<br />

Honeybees have numerous olfactory sensory hairs that make up their<br />

sensory epithelium, which is distributed along the length of the<br />

antennae. The antennae are shaped as tubular structures that can be<br />

placed in different orientations and moved with low frequency through<br />

air. In general, antennal sensilla increase in number from proximal to<br />

distal. Because of this uneven distribution of sensors and the way in<br />

which the shape of the antennae interacts physically with the<br />

environment, antennal movements have important consequences <strong>for</strong><br />

odorant molecule capture and, there<strong>for</strong>e, perception. Antennal<br />

movements were videotaped during presentation of two odorants<br />

(geraniol and 1-hexanol) and a control blank. Bee responses were<br />

recorded both be<strong>for</strong>e and after being trained to associate a food reward<br />

with an odorant, as demonstrated by a proboscis extension response.<br />

The movements of the antennae were analyzed in three-dimensional<br />

space by digitizing the base, elbow, and distal tip of each antenna.<br />

Antennal movements increased in response to odorants. Antennal<br />

movements were not simple oscillations in a plane but were<br />

complicated excursions in three-dimensional space. Antennae were<br />

oriented in an anterior direction more noticably in response to geraniol.<br />

Left and right antennae are moved independently of each other. These<br />

increased movements will increase the volume of air sampled by the<br />

antennae and a potential increase in odorant capture rate from the<br />

moving air.<br />

This word was supported by an award from NIH-NCRR to BHS (9<br />

R01 RR14166) and from NSF to CL (IBN-9984475).<br />

68<br />

263 Poster [ ] Chemical Ecology<br />

OLFACTORY-MEDIATED SEARCH BEHAVIORS OF<br />

MIGRATORY SEA LAMPREYS SEEKING PHEROMONE-<br />

LADEN SPAWNING STREAMS IN THE GREAT LAKES<br />

Vrieze L.A. 1, Sorensen P.W. 1 1Department of Fisheries, Wildlife and<br />

Conservation Biology, University of Minnesota, St. Paul, MN<br />

The sea lamprey (Petromyzon marinus) spends its larval life in<br />

streams, enters large lakes or oceans as a parasitic juvenile, and<br />

eventually returns to streams as an adult to reproduce. Laboratory<br />

studies strongly suggest that lampreys locate suitable spawning streams<br />

using a bile acid-related migratory pheromone released by larval<br />

conspecifics (see Fine & Sorensen, this symposium). This study<br />

examined olfactory-mediated search behaviors of lampreys as they<br />

searched <strong>for</strong> pheromone-laden streams emptying into Lake Huron. Both<br />

olfactory-occluded and untreated lampreys implanted with acoustic<br />

telemetry tags were tracked from a GPS-equipped boat and their<br />

movements related to stream water concentration (as measured by<br />

conductivity). Outside the apparent influence of stream water, lampreys<br />

(N=12) swam rapidly (1.5 km/hr) and continuously on relatively<br />

straight bearings (straightness index = 0.78) while per<strong>for</strong>ming repeated<br />

vertical excursions through the water column. Occluded and intact<br />

animals swam with different compass bearings, perhaps due to differing<br />

responses to lake currents/odors. Upon encountering river plumes<br />

(N=10), lampreys with an intact olfactory sense commenced circling<br />

(straightness index = 0.42). The speed and success of locating the<br />

stream mouth was correlated with the thoroughness of mixing of the<br />

stream and lake waters. These behaviors, together with responses to<br />

stream waters previously documented in laboratory mazes, suggest the<br />

use of klinotactic chemo-orientation in stream finding. Funded by the<br />

Great Lakes Fisheries Commission.<br />

264 Poster [ ] Chemical Ecology<br />

CHEMICAL FRACTIONATION DEMONSTRATES THAT THE<br />

SEA LAMPREY MIGRATORY PHEROMONE IS COMPRISED<br />

OF SEVERAL BILE ACID-LIKE COMPOUNDS<br />

Fine J.M. 1, Sorensen P.W. 1 1Department of Fisheries, Wildlife and<br />

Conservation Biology, University of Minnesota, St Paul, MN<br />

The sea lamprey (Petromyzon marinus) starts its life in freshwater<br />

streams which it then leaves to parasitize lake/oceanic fishes be<strong>for</strong>e<br />

eventually re-entering streams to spawn. Laboratory and field studies<br />

have shown that adult lampreys locate spawning streams using a<br />

pheromone released by stream-resident larval lampreys (see Vrieze and<br />

Sorensen, this symposium). Initial biochemical characterization of this<br />

cue found it to contain the sulfated bile acid petromyzonol sulfate (PS)<br />

which adult lampreys detect at 10 -12 Molar (M). Using a combination<br />

of HPLC fractionation, olfactory recording, behavioral assays, and mass<br />

spectrometry we have recently isolated two additional compounds from<br />

larval holding water that have pheromonal activity. The most important<br />

has a molecular weight of 704, co-elutes with PS by HPLC, and is<br />

behaviorally attractive at concentrations below 10 -14 M, a record <strong>for</strong><br />

fish. The second has a molecular weight of 590 and is less potent. In a<br />

two-choice preference maze, adult lampreys did not distinguish<br />

between larval water and a mixture comprised of PS and these two<br />

compounds, demonstrating that this mixture constitutes the majority of<br />

the pheromone. It is as yet unclear whether the lamprey has evolved to<br />

respond to multiple compounds released by larvae to increase their<br />

sensitivity to larval odor or to discern it more specifically. Ef<strong>for</strong>ts are<br />

presently underway to elucidate the structures of the unknown<br />

compounds. Funded by the Great Lakes Fishery Commission.


265 Slide [ ] Chemical Ecology<br />

LARVAL REEF FISH DISCRIMINATE BETWEEN REEF<br />

ODORS AND MAY USE THIS IN RECRUITMENT.<br />

Atema J. 1, Gerlach G. 2, Kings<strong>for</strong>d M. 3 1Marine Program, Boston<br />

University, Woods Hole, MA; 2Marine Resources, Marine Biological<br />

Laboratory, Woods Hole, MA; 3James Cook University, Townsville,<br />

Queensland, Australia<br />

Larval reef fishes typically have an early pelagic phase from which<br />

they must return back to the reef environment to survive. The<br />

discrepancy between passive dispersal models and actual recruitment<br />

data suggests that these small animals participate actively in the<br />

recruitment process, but the behavioral mechanisms are not clear. The<br />

ontogeny of swimming efficiency of several species is now known;<br />

sensory capabilities remain poorly understood. We have been pursuing<br />

the olfactory hypothesis that early imprinting on the home reef odor<br />

allows them to remain near the (natal) reef and perhaps may guide their<br />

return. We have now demonstrated <strong>for</strong> different species that at the stage<br />

that they return to the reef they can discriminate between the odor of<br />

different reefs and among 5 choices prefer the odor of the reef to which<br />

they returned. We are now using genetic markers to see if this sensory<br />

capability has led to strong homing that is reflected in population<br />

genetic substructuring at the scale of single reefs separated by kilometer<br />

distances with and without current barriers.<br />

266 Poster [ ] Chemical Ecology<br />

FRUIT ODOR DISCRIMINATION AND HOST RACE<br />

FORMATION IN RHAGOLETIS FRUIT FLIES<br />

Linn C. 1, Nojima S. 1, Roelofs W. 1 1Entomology, Cornell University,<br />

Geneva, NY<br />

Rhagoletis pomonella is a model <strong>for</strong> incipient sympatric speciation<br />

(divergence without geographic isolation) via host plant shifts. We are<br />

testing the hypothesis that R. pomonella flies originating from<br />

hawthorn, apple, and flowering dogwood fruit use fruit volatiles to<br />

distinguish among their respective host plants. Solid phase<br />

microextraction combined with gas chromatography and<br />

electroantennogram detection were used to identify unique blends of<br />

volatiles from each fruit type. In flight tunnel assays flies preferentially<br />

flew upwind (>70%) to the volatile blend from their natal host. Field<br />

tests also showed that over the fruiting season significantly more flies<br />

were captured with natal fruit volatiles. Because R. pomonella<br />

rendezvous on or near the unabscised fruit of their hosts to mate, the<br />

behavioral preference <strong>for</strong> natal fruit odor translates directly into<br />

premating reproductive isolation between the fly races. To explore the<br />

genetic basis of the trait crosses were produced between the different<br />

fly populations. Flight tunnel tests showed that hybrid flies have<br />

significantly reduced response levels to natal or combined volatile<br />

blends, with only 40% of the flies exhibiting upwind flight, and only<br />

with a concentration ten times that used with parent flies. Flies<br />

generated from F2 crosses exhibited response profiles similar to F1<br />

hybrids (60%) or to the parent populations (40%). The behavioral data<br />

lay the foundation <strong>for</strong> QTL analysis. Supported by NSF DEB-9977011.<br />

69<br />

267 Poster [ ] Chemical Ecology<br />

CO2 IS INVOLVED IN THE OVIPOSITION BEHAVIOR OF<br />

MANDUCA MOTHS<br />

Guerenstein P.G. 1, Abrell L. 2, Mechaber W.L. 1, Stange G. 3, Hildebrand<br />

J.G. 1 1ARL Division of Neurobiology, Univ. of Arizona, Tucson, AZ;<br />

2Biosphere 2 Chemistry Unit, Columbia Univ., Oracle, AZ; 3Biological<br />

<strong>Sciences</strong>, Australian National Univ., Canberra, Australia<br />

Moths detect CO in the environment, but the role of CO in the<br />

2 2<br />

biology of these insects is not well understood. It has been suggested<br />

that CO plays a role in the oviposition of the moth Cactoblastis<br />

2<br />

cactorum and in the nectar <strong>for</strong>aging of Manduca sexta. We asked if<br />

CO also plays a role in oviposition in Manduca. Two groups of plants,<br />

2<br />

each surrounded by a ring of ductwork, were placed inside a flight cage.<br />

Fans pumped ambient air into the ducts in which small holes provided<br />

<strong>for</strong> gas outflow. CO was injected at the inlet <strong>for</strong> one group of plants,<br />

2<br />

thus generating an artificial plume of high CO around that group.<br />

2<br />

Female Manduca were released into the cage singly. We measured<br />

number of eggs oviposited. We found that, as in Cactoblastis, females<br />

preferred the control plants (no CO added), possibly because natural<br />

2<br />

CO fluctuations generated by the test plants were masked by an<br />

2<br />

artificially high CO plume. Thus, CO affects the oviposition behavior<br />

2 2<br />

of Manduca. We did not find a dose effect, suggesting that the moths´<br />

response was saturated at the CO levels used. The response of<br />

2<br />

Manduca was lower than that reported <strong>for</strong> Cactoblastis. Because<br />

Cactoblastis oviposits on (CAM) plants that generate sinks of CO on 2<br />

their surface while Manduca oviposits on plants that are sources of<br />

CO , we suggest that the expected rise in global ambient CO levels<br />

2 2<br />

will affect more strongly moths that rely on CO sinks than those that<br />

2<br />

rely on CO sources. [Supported by NSF (IBN-0213032), Columbia<br />

2<br />

Chemistry, Biosphere 2]<br />

268 Poster [ ] Chemical Ecology<br />

DEVELOPMENTAL EXPRESSION AND TISSUE<br />

DISTRIBUTION OF AN ODORANT-BINDING PROTEIN (OBP)<br />

IN THE MALE YELLOW FEVER MOSQUITO AEDES<br />

AEGYPTI<br />

Bohbot J. 1, Vogt R. 1 1Biological <strong>Sciences</strong>, University of South<br />

Carolina, Columbia, SC<br />

Major concerns of an adult male mosquito are finding mates and food<br />

sources. While most studies concentrate on female specific olfactory<br />

processes such as seeking a host <strong>for</strong> a suitable blood meal, little is<br />

known regarding other olfactory based behaviors, including those of the<br />

male. Even less is known at the molecular level. We have cloned and<br />

characterized an OBP from a male antennal cDNA library. The Aaeg-<br />

OBP sequence is most similar to members of the Drosophila and<br />

Anopheles OBP family. Due to the limited amount of material, we used<br />

the quantitative Real-Time PCR method to measure expression levels<br />

among different tissues and at different developmental times in both<br />

males and females. Aaeg-OBP expression is adult and male specific.<br />

Moreover, its expression is restricted to appendages known to carry<br />

chemosensory organs, including antennae and wings. The expression of<br />

this gene in adult but not larval males suggests the protein functions in a<br />

male specific adult context, mediating chemosensory signals important<br />

to behaviors that are more relevant to the males than to the females. The<br />

identification of such neural pathways significantly expands our<br />

opportunity <strong>for</strong> developing strategies which disrupt chemosensory<br />

mosquito behavior and thus mitigate human-mosquito interactions.


269 Poster [ ] Chemical Ecology<br />

MECHANISMS OF ACTION OF DEFENSIVE SECRETIONS OF<br />

THE SEA HARE APLYSIA CALIFORNICA AGAINST THE<br />

SPINY LOBSTER PANULIRUS INTERRUPTUS<br />

Shabani S. 1, Derby C.D. 1, Kicklighter C. 1, Johnson P. 1 1Biology,<br />

Georgia State University, Atlanta, GA<br />

Active chemical defenses of Aplysia, which include secretions from<br />

the opaline and ink glands, deter attacks by potential predators. The<br />

mechanism of action of these secretions differs among predators.<br />

Against the sea anemone Anthopleura sola, ink from Aplysia<br />

cali<strong>for</strong>nica contains a protein that is aversive and cytolytic (Kicklighter<br />

et al., 2004 AChemS poster). Against the spiny lobster Panulirus<br />

interruptus, sea hares that release secretions upon attack are more likely<br />

to be dropped, allowing sea hares to escape. Ink and opaline appear to<br />

protect sea hares through a combination of phagomimicry, chemical<br />

confusion, and aversion. Phagomimicry, in which predators are<br />

deceived into attending to a false food stimulus from the secretions, and<br />

chemical confusion occur because the defensive secretions contain<br />

millimolar concentrations of amino acids and coat the sensory organs of<br />

lobsters. These levels of amino acids in ink and opaline are enormously<br />

stimulatory to the lobster´s chemosensory cells, and the stickiness likely<br />

causes a persistent stimulation. The aversive effect against lobsters is<br />

mediated by components in opaline. In addition, ink and opaline are<br />

highly acidic (4.9 and 5.8 respectively), suggesting that pH may have<br />

either a direct or indirect effect on the secretions´ efficacy. We are<br />

currently examining these pH effects in behavioral and<br />

electrophysiological assays. Supported by NSF IBN-0324435<br />

270 Poster [ ] Chemical Ecology<br />

PROTEIN-MEDIATED DEFENSE IN APLYSIA CALIFORNICA<br />

AGAINST THE PREDATORY ANEMONE ANTHOPLEURA<br />

SOLA<br />

Kicklighter C. 1, Johnson P. 1, Yang H. 1, Tai P. 1, Derby C. 1 1Biology,<br />

Georgia State University, Atlanta, GA<br />

The sea hare Aplysia cali<strong>for</strong>nica defends itself from predators with<br />

two secretions, ink and opaline, which affect predators differently.<br />

Against the spiny lobster Panulirus interruptus ink is attractive while<br />

opaline is aversive (Shabani et al., 2004 AChemS poster). Conversely,<br />

against the predatory sea anemone Anthopleura sola, ink is aversive,<br />

causing tentacles to shrivel and the gastrovascular cavity to evert.<br />

Opaline appears to stimulate feeding, as it elicits tentacle movement to<br />

the mouth, which opens. Our investigations indicate that ink contains<br />

one dominant protein, “escapin”, which occurs at a concentration of<br />

0.025 mg/ml of ink. When applied to anemone tentacles, escapin elicits<br />

cell lysis, suggesting that escapin may be responsible <strong>for</strong> ink´s<br />

aversiveness to sea anemones. To further pursue escapin´s defensive<br />

role <strong>for</strong> Aplysia, we are using the technique RNA interference to knock<br />

down production of escapin. Ink produced by Aplysia injected with<br />

double-stranded RNA elicits less tentacle shriveling than ink produced<br />

by control Aplysia. We are currently using this technique to assess the<br />

survival of sea hares that can and cannot produce ink containing escapin<br />

against sea anemones.<br />

Supported by NSF IBN-0324435<br />

70<br />

271 Poster [ ] Chemical Ecology<br />

PREDATOR ODORS AND REPRODUCTION IN HOUSE<br />

MOUSE UNDER LABORATORY AND SEMI-NATURAL<br />

CONDITIONS<br />

Voznessenskaya V. 1, Naidenko S. 1, Dulchenko N. 1, Clark L. 2 1Institute<br />

of Ecology & Evolution RAS, Moscow, Russia; 2Repellents, National<br />

Wildlife Research Center, Fort Collins, CO<br />

We examined the influence of predator odors (Felis catus) on<br />

reproductive output of House Mouse (Mus musculus) under laboratory<br />

and semi-natural conditions. Laboratory naive animals responded to<br />

predator chemical cues either with block of pregnancy or reduced litter<br />

size and skewed sex ratio. Under laboratory conditions block of<br />

pregnancy in experimental group was observed twice higher then in<br />

both of the control groups (p< 0.001). In enclosures 30% of females<br />

were not pregnant (season of May-September 2002-2003) when cat<br />

urine was applied onto the bedding each other day. At the same time in<br />

control enclosure 92 % (n=38) of females were pregnant. The total<br />

number of offspring in control enclosure also was significantly higher<br />

(p< 0.001) then in experimental one. We observed seasonal changes in<br />

sensitivity of mice to predator chemical cues. It is noteworthy that<br />

suppression of rodent reproduction also occurred when rodents were<br />

exposed to urine of conspecifics housed under high population<br />

densities. The fact that mice respond to certain chemical signals in<br />

predator urine in similar fashion may be <strong>for</strong>tuitous, and may have more<br />

to do with the coincidence that the urine contain similar chemical cues<br />

resulting from protein digestion in carnivores and protein catabolism in<br />

nutritionally deprived rodents rather than specific predator-prey<br />

adaptations.<br />

Supported by RFBR & Presidium RAS “Biological Resources” #<br />

3.1.1.<br />

272 Slide [ ] Chemical Ecology<br />

MANUFACTURE AND TESTING OF CHEMICAL-SIGNAL-<br />

ENHANCED DEVICES FOR DETERRING CROP-RAIDING<br />

ELEPHANTS<br />

Rasmussen L.E. 1, Riddle S.W. 2, Roeder H. 2 1OGI School of Science &<br />

Engineering, OHSU, Beaverton, OR; 2Riddle's Elephant & Wildlife<br />

Sanctuary, Greenbrier, AR<br />

Human-elephant conflict is a significant economic and ecological<br />

problem in southeast Asia, escalating as human populations continue to<br />

expand and the <strong>for</strong>est habitat <strong>for</strong> elephants is destroyed. In their native<br />

society Asian elephants utilize two pheromones that elicit well-defined<br />

behavioral responses and a variety of chemical signals that influence<br />

other behaviors including movement and choices. Utilizing this<br />

in<strong>for</strong>mation combined with several years of testing various antifeedants<br />

and deterrents, and detailed first-hand knowledge of the<br />

walking behavior of elephants, Riddle and Rasmussen invented a<br />

mechanical device, enhanced with chemical signal dispersion units.<br />

Recently we built twelve such devices at a village site in southern India<br />

and tested their efficacy during harvest season, the period of maximum<br />

crop-raiding by elephants. As crop-raiding in this region occurs<br />

primarily nocturnally, observations were conducted using night vision<br />

recording equipment. Results demonstrate that 1. the manufacturing<br />

process was cost effective 2. the devices were efficacious as deterrents<br />

3. new ambulatory capabilities and high risk behaviors by wild male<br />

elephants were revealed. We present this study in the context of the<br />

relevance of basic behavioral and chemical signal research toward<br />

practical economic use. Supported by USFW grants # 98210-G091to<br />

LELR and # 98210-3-G648 to LELR and SR.


273 Poster [ ] Accessory Olfactory System<br />

MODIFICATION OF ODOR INVESTIGATION BY FEMALE<br />

OPOSSUMS (MONODELPHIS DOMESTICA) AFTER<br />

ACCESSORY OLFACTORY BULB ABLATION<br />

Zuri I. 1, Halpern M. 1 1Anatomy & Cell Biology, Downstate Medical<br />

Center, Brooklyn, NY<br />

Monodelphis domestica are South American marsupials that<br />

commonly scent-mark their environment. We have previously shown<br />

that female opossums discriminate between conspecific odors from<br />

different sources. The purpose of the present study was to determine if<br />

the vomeronasal system is important <strong>for</strong> females to discriminate these<br />

odors.<br />

Investigation of conspecific odors and distilled water was tested in<br />

12 female opossums in a 2-choice paradigm and the time spent in snout<br />

contact with each stimulus was analyzed from video recordings.<br />

Females were tested be<strong>for</strong>e treatments and after control treatment<br />

(anaesthesia and partial elctrolytic accessory olfactory bulb lesions<br />

(AOBP, N=6), or anaesthesia followed by complete AOB ablation<br />

(AOBX, N=6)). Neither anesthesia nor AOBP had an effect on odor<br />

investigation.<br />

AOBX resulted in a significant reduction of investigation of certain<br />

odors, but only when those odors were paired with certain other odors.<br />

For example, when male suprasternal gland (SG) odor was paired with<br />

urine of the same stimulus male, there was a significant decrease in the<br />

time spent investigating SG odors following AOBX. However, AOBX<br />

did not modify the investigation of male SG when it was paired with the<br />

same-male sub-mandibular odors.<br />

Our data suggest that the accessory olfactory system may be<br />

important <strong>for</strong> the investigation of conspecific odors by female<br />

opossums, but it does not appear to be essential <strong>for</strong> such odor<br />

discrimination.<br />

Supported by NIH Grant DC02745.<br />

274 Poster [ ] Accessory Olfactory System<br />

TWO POPULATIONS OF GRANULE CELLS IN THE<br />

ACCESSORY OLFACTORY BULB OF THE OPOSSUM,<br />

MONDELPHIS DOMESTICA<br />

Jia C. 1, Halpern M. 2 1Anatomy and Cell Bliology, Downstate Medical<br />

Center, Brooklyn, NY; 2Anatomy & Cell Biology, Downstate Medical<br />

Center, Brooklyn, NY<br />

In the accessory olfactory bulb (AOB) the principal neurons,<br />

mitral/tufted cells, receive synaptic inputs from vomeronasal receptor<br />

cell axons. Mitral/tufted cell activity is modulated by feedback control<br />

of interneurons. Granule cells are the major type of inhibitory<br />

interneurons. In this study, we used immunocytochemical technique to<br />

analyze the expression of the calcium binding proteins calretinin and<br />

calbindin D28k in the granule cell layer of the opossum AOB. We<br />

identified two separate populations of granule cells in the granule cell<br />

layer. One population expresses only calretinin, and the other expresses<br />

only calbindin D28k. Their somatic distribution in the granule cell layer<br />

was intermingled and their dendritic projection into the external<br />

plexi<strong>for</strong>m layer was also similar. No such differential expression of the<br />

calcium binding proteins was detected in the mouse or rat AOB, or in<br />

the main olfactory bulbs of the rat, mouse and opossum using the same<br />

antibodies. This observation suggests that in the opossum AOB the<br />

granule cell population is not homogenous.<br />

Supported by NIDCD grant #DC02745<br />

71<br />

275 Poster [ ] Accessory Olfactory System<br />

VOMERONASAL AND OLFACTORY CONVERGENCE IN<br />

MEDIAL AMYGDALA<br />

Case G.R. 1, Meredith M. 2 1Neurosciences Program, Florida State<br />

University, Tallahassee, FL; 2Neuroscience Program, Florida State<br />

University, Tallahassee, FL<br />

Conspecific chemosensory signals communicate social in<strong>for</strong>mation<br />

in hamsters, and are distinguished from heterospecific, socially nonrelevant,<br />

stimuli by Fos responses within medial amygdala (Westberry<br />

and Meredith AChemS 2003). Signals stimulate the vomeronasal organ<br />

(VNO) or main olfactory system, activating medial amygdala via<br />

accessory and main olfactory bulbs (AOB, MOB). We used electrical<br />

stimulation of the VNO and MOB in anesthetized hamsters to show that<br />

both systems have input to the medial amygdala. MOB electrodes were<br />

in the rostro-lateral bulb to avoid driving VNO nerves. Simultaneous<br />

recordings were made with 4 electrodes inserted into a 3 x 8 grid<br />

pattern (200 um spacing) covering most of the anterior (MeA) and<br />

posterior (MeP) medial amygdala. Robust field potentials appeared<br />

throughout with moderate to high stimulus intensity (400-1000 uA,<br />

500uS) and a few single units were driven by both systems. Full<br />

mapping of MeA and MeP was obtained in 5 hamsters and partial<br />

mapping of MeP in 2 others, as verified by histology. Olfactory input<br />

clearly has access to medial amygdala even though it was not<br />

necessary, in the previous Fos studies, <strong>for</strong> stimulus categorization by<br />

MeA/MeP. Electrical VNO stimulation in awake hamsters activated Fos<br />

expression in MeA but not MeP (as <strong>for</strong> heterospecific chemosensory<br />

stimulation). Here we found approx equal efficiency <strong>for</strong> short-latency<br />

VNO driven activation of MeA and MeP, possibly due to higher<br />

stimulus levels than used in awake animals (150uA), or to the<br />

cumulative effect of activation on Fos expression over 15 min. These<br />

hypotheses will be tested in future experiments. Support by NIDCD<br />

grant DC005813<br />

276 Poster [ ] Accessory Olfactory System<br />

CATEGORIZATION OF CHEMOSENSORY INPUT IN<br />

MEDIAL AMYGDALA REQUIRES VOMERONASAL INPUT IN<br />

BOTH SEXUALLY NAÏVE AND EXPERIENCED MALE<br />

HAMSTERS.<br />

Westberry J. 1, Samuelsen C.L. 2, Meredith M. 1 1Program in<br />

Neuroscience, Florida State University, Tallahassee, FL;<br />

2Neuroscience, Florida State University, Tallahassee, FL<br />

In male hamsters, medial amygdala responds categorically to<br />

chemosensory input based on the species and social relevance of a<br />

stimulus. Using immediate early gene (IEG) expression, we have<br />

demonstrated that the anterior medial amygdala (MeA) responds to both<br />

conspecific and heterospecific stimuli, but posterior medial amygdala<br />

(MeP) responds only to conspecific (socially-relevant) stimuli. Another<br />

part of the amygdala, the largely GABAergic intercalated nucleus<br />

(ICN) was activated when MeP was not activated with heterospecific<br />

stimuli, suggesting inhibition of MeP by ICN. This categorization<br />

appears to be hard-wired. In sexually-naïve males, lesions of the main<br />

olfactory epithelium (OLFX) did not change the pattern of IEG<br />

activation in the MeA or MeP elicited by conspecific and heterospecific<br />

stimuli, while removal of the VNO (VNX) eliminated categorization<br />

and most of the responses in the MeA and MeP. Sexually-experienced<br />

males can use main olfactory input to compensate <strong>for</strong> lack of VNO<br />

input. Here we investigated activation and categorization in MeA and<br />

MeP in experienced-VNX males. Preliminary data indicates that<br />

experienced-VNX males show activation of MeA and MeP with both<br />

conspecific and heterospecific stimuli, indicating a lack of<br />

categorization of these stimuli. Without VNO input, there was never<br />

activation of ICN with either class of stimuli. There were no significant<br />

differences in attention to the stimuli (on swabs) that could account <strong>for</strong><br />

differences in amygdala FRAs expression between groups. Supported<br />

by DC-005813 from NIDCD.


277 Poster [ ] Accessory Olfactory System<br />

CORTICAL RESPONSE TO ANDROSTADIENONE WITH OR<br />

WITHOUT FUNCTIONAL OCCLUSION OF THE<br />

VOMERONASAL DUCT - A FUNCTIONAL MAGNETIC<br />

RESONANCE IMAGING (FMRI) STUDY<br />

Gerber J.C. 1, Lundstrom J.N. 2, Frasnelli J. 3, Knecht M. 3, Olsson M. 2,<br />

Hummel T. 3 1Neuroradiology, Dresden University, Dresden, Germany;<br />

2Psychology, Uppsala University, Uppsala, Sweden;<br />

3Otorhinolaryngology, Dresden University, Dresden, Germany<br />

There has been controversy as to whether the vomeronasal duct<br />

(VND) in humans is mediating any sensory in<strong>for</strong>mation or is merely a<br />

nonfunctional vestige. This study assessed the cortical responses<br />

following stimulation with the putative pheromone Androstadienone<br />

(AND) in women, with and without coverage of the VND and using<br />

AND in concentrations beyond olfactory threshold. We hypothesised<br />

that activation is independent of the VND´s status. Following detailed<br />

examination including tests <strong>for</strong> olfactory functions, 16 women (age<br />

range 21 to 27 years), fertile, and all presenting with a VND underwent<br />

fMRI. The odour stimuli were AND (3 mol/l) and phenylethyl alcohol<br />

(PEA) (pseudo randomised block design, 30s on, 30s off; 1s stimulus<br />

duration, 3s interstimulus interval). Subjects did not know about the<br />

status of the VND´s coverage. Imaging data were analysed with SPM2.<br />

In a group random effects analysis, coverage of the VND did not show<br />

significant influence on cortical activation. On first sight, AND<br />

produced an activation pattern similar to that of the control odour PEA.<br />

When further assessing the differences (AND vs. PEA, group random<br />

effects analysis), significant subcallosal activation extending to the<br />

septal region was detected. In summary, access to the VND does not<br />

seem to play a major role in the perception of AND odor. The<br />

subcallosal and septal activation specific to AND is in line with<br />

previous findings with PET. Support: DFG HU441-2, HSFR:F0868<br />

278 Poster [ ] Accessory Olfactory System<br />

CHARACTERISTICS OF GENERAL AND SPECIFIC<br />

CHEMOSENSORY RESPONSES IN THE SNAKE ACCESSORY<br />

OLFACTORY BULB<br />

Cinelli A. 1, Li C. 2, Wang D. 3, Liu W. 4, Chen P. 3, Halpern M. 5<br />

1Downstate Medical Center, Brooklyn, NY; 2Anatomy & Neurobiology,<br />

University of Tennessee, Memphis, TN; 3Biochemistry, Downstate<br />

Medical Center, Brooklyn, NY; 4Anatomy and Cell Biology, Downstate<br />

Medicall Center, Brooklyn, NY; 5Anatomy & Cell Biology, Downstate<br />

Medical Center, Brooklyn, NY<br />

Converging evidence indicates that the main olfactory system detects<br />

general odors using a combinatorial coding strategy. Currently, the<br />

coding strategy in the vomeronasal (VN) system is unknown, but this<br />

system detects highly selective chemosignals, and probably depends on<br />

the activation of highly tuned VN receptor neurons. In snakes, the VN<br />

system is critical <strong>for</strong> detecting chemosignals emitted by conspecifics<br />

and prey, but also responds to general odors and amino acids when they<br />

are delivered as liquid stimuli. Using voltage sensitive dyes and<br />

selective chemoattractants, we studied the distribution of neural<br />

responses evoked by electrical, general and specific chemosensory<br />

stimuli in the snake accessory olfactory bulb (AOB). All stimulus types<br />

evoke non-homogeneously distributed patterns within and throughout<br />

the layers, and seemed to be organized in clusters. The spatial<br />

distribution of responses to prey chemicals suggest the presence of a<br />

zone of preferential activation located in the posterior AOB. Changes in<br />

the spatial distribution of activity occurred over time, suggesting a<br />

stimulus-specific dynamic organization of responses. Foci of activity<br />

also exhibited dissimilar thresholds. These results suggest that sensory<br />

discrimination in the snake AOB is based on a combinatorial coding<br />

scheme.<br />

Supported by NIH DC 03735<br />

72<br />

279 Slide [ ] Olfactory Sensory Neuron Physiology<br />

CHLORIDE HOMEOSTASIS IN MOUSE ORNS<br />

Reisert J. 1, Yau K. 1, Bradley J. 1 1Department of Neuroscience/HHMI,<br />

Johns Hopkins University, Baltimore, MD<br />

Olfactory transduction was seen as a system that worked analogously<br />

to vertebrate phototransduction: Stimulation of G protein-coupled<br />

receptors triggers an electrical response by activating CNG channels.<br />

This view was altered when Kleene & Gesteland reported the presence<br />

of a Ca2+-activated Cl channel on olfactory cilia. Odor stimulation<br />

increases the ciliary Ca2+ concentration by opening CNG channels,<br />

generating a secondary Cl current. This current is inward, there<strong>for</strong>e<br />

excitatory, implying a relatively high internal chloride concentration.<br />

80% of the odor-induced current in rodents and 36-65 % in amphibians<br />

is carried by chloride. Two functions <strong>for</strong> this unusual Cl current have<br />

been proposed. First, it provides a nonlinear, low-noise amplification of<br />

the CNG current. Second, it reduces the dependency of the receptor<br />

current on external mucosal Na+, which might vary depending on the<br />

environment, a situation more relevant <strong>for</strong> amphibians and fish than <strong>for</strong><br />

mammals. Many of the characteristics of the odor-induced response<br />

there<strong>for</strong>e depend on the interplay between the CNG channel (the Ca2+<br />

source) and the Cl channel, and their regulation. Furthermore, since<br />

spike frequency is a function of the overall receptor current, the<br />

interaction of these two channels will affect action-potential generation.<br />

We present molecular and electrophysiological evidence <strong>for</strong> a<br />

mechanism by which internal Cl– concentration is maintained such that<br />

its reversal potential is positive relative to the resting membrane<br />

potential in mouse ORNs.<br />

280 Poster [ ] Olfactory Sensory Neuron Physiology<br />

CHLORIDE HOMEOSTASIS IN MOUSE OLFACTORY<br />

NEURONS<br />

Delay R. 1, Verret T. 1, Gorman R. 1 1Biology, University of Vermont,<br />

Burlington, VT<br />

Intracellular Cl, [Cl]i, is important in determining the ultimate<br />

response of olfactory sensory neurons (OSNs) exposed to odor.<br />

Although secondary to the non-selective cation current, the Cadependent<br />

Cl current is primarily responsible <strong>for</strong> the generator<br />

potential. Thus, Cl homeostasis is important to odor responses. We used<br />

MEQ, a chloride sensitive fluorescent dye, to measure [Cl]i across a<br />

population of isolated OSNs. Since MEQ is a non-ratiometric dye, a set<br />

of Cl bath standards was used to establish the initial level of internal Cl.<br />

Our results indicated a wide range of [Cl]i exist in isolated OSNs<br />

(range: 20-145 mM; mean: 62 mM). As a control, [Cl]i measured in<br />

erythrocytes had a mean of 55 +/- 5 mM, similar to other reports <strong>for</strong><br />

RBCs. Although the population of isolated OSNs displayed a wide<br />

range of [Cl]i, in any individual OSN [Cl]i was stable during the<br />

recording period (up to two hours). The only changes observed in [Cl]i<br />

were due to stimulation or altering bath conditions to affect Cl<br />

cotransporters. [Cl]i of OSNs responded in a dynamic fashion, changing<br />

in response to odor stimulation or by increasing [cAMP]i. With [Cl]i<br />

unusually high in some olfactory neurons, there must be an active<br />

transporter or exchanger(s) that maintains this gradient. Several<br />

different transporters have been reported to carry Cl in neurons. We<br />

focused on NKCC1 and KCC2 since these Cl cotransporters have been<br />

shown to be critical to [Cl]i in developing neurons. Drugs were used to<br />

selectively inhibit either NKCC1 or KCC2. Data from these<br />

experiments, imaging and per<strong>for</strong>ated patch clamp, suggest these<br />

transporters are functional in OSNs and work to set [Cl]i.


281 Poster [ ] Olfactory Sensory Neuron Physiology<br />

EXPRESSION OF CL- COTRANSPORTERS IN MOUSE<br />

OLFACTORY NEURONS.<br />

Schannen A.P. 1, Delay R.J. 2 1Biology, University of Vermont,<br />

Burlington, VT; 2Biology Department, University of Vermont,<br />

Burlington, VT<br />

Chloride fluxes and intracellular Cl, [Cl]i, are important in<br />

determining the ultimate response of olfactory sensory neurons (OSN)<br />

exposed to odor. In some OSNs, Cl can contribute most of the odorelicited<br />

generator current, giving it a vital role in olfactory signaling.<br />

Populations of isolated mouse OSNs have been shown to express a<br />

wide range [Cl]i. To determine what Cl transporters might be involved<br />

in this variance we characterized the expression of two chloride<br />

cotransporters in mouse OSNs: iso<strong>for</strong>m 1 of the Na/K/2Cl cotransporter<br />

(NKCC1) and iso<strong>for</strong>m 2 of the K/Cl cotransporter (KCC2). NKCC1<br />

moves one Na ion, one K ion and 2 Cl ions into the cell. Conversely,<br />

KCC2 moves 1 K and 1 Cl ion out of the cell. Since these are<br />

cotransporters, if any one of ions being moved across the membrane is<br />

not present, no movement occurs. Western blots of nasal epithelium and<br />

immunohistochemistry of sectioned tissues, showed that both KCC2<br />

and NKCC1 were present in mouse olfactory neurons and both<br />

transporters appeared to be located in the apical region of OSNs. To<br />

further characterize the expression patterns of KCC2 and NKCC1 we<br />

examined isolated cells using deconvolution microscopy. OSNs showed<br />

variable expression of KCC2 and NKCC1. Some OSNs displayed<br />

intense labeling <strong>for</strong> NKCC1 and KCC2 in their dendritic region, while<br />

others express these cotransporters at lower levels throughout the entire<br />

cell. A few OSNs showed only faint expression of either cotransporter.<br />

We have shown that NKCC1 and KCC2 are present in mouse OSNs,<br />

although expression is not uni<strong>for</strong>m across the population of OSNs.<br />

Supported by NIH grants P20RR16435 & P20RR16462<br />

282 Poster [ ] Olfactory Sensory Neuron Physiology<br />

PENDRIN, A CHLORIDE TRANSPORTER, IS EXPRESSED IN<br />

OLFACTORY RECEPTOR NEURONS<br />

Kleene N.K. 1, Zhang J. 1, Pixley S.K. 1, Soleimani M. 2, Kleene S.J. 1 1Cell<br />

Biology, Neurobiology & Anatomy, University of Cincinnati,<br />

Cincinnati, OH; 2Internal Medicine, University of Cincinnati,<br />

Cincinnati, OH<br />

During odorant transduction, a calcium-activated chloride current<br />

contributes to the depolarization of olfactory receptor neurons (ORNs).<br />

The purpose of this study is to identify anion exchangers that regulate<br />

internal chloride concentration in ORNs. Toward this end, reverse<br />

transcriptase-polymerase chain reaction was per<strong>for</strong>med on RNA<br />

isolated from murine nasal mucosa using primers specific <strong>for</strong> members<br />

of the SLC26 family, a highly conserved group of anion exchangers<br />

that transport chloride. Our results identified the expression of pendrin<br />

in murine nasal mucosa. Pendrin is an ion transporter that exchanges<br />

chloride <strong>for</strong> bicarbonate, hydroxide, iodide, or <strong>for</strong>mate but not sulfate.<br />

It also exchanges chloride <strong>for</strong> fructose and mannose. Pendrin is<br />

abundantly expressed in a distinct subpopulation of cells in the thyroid,<br />

inner ear, and kidney. Mutations in pendrin cause Pendred´s syndrome,<br />

a hereditary disorder characterized by deafness and goiter. Using<br />

Northern blotting, we confirmed that pendrin mRNA is expressed in<br />

nasal mucosa. Immunohistochemical staining with specific antibodies<br />

indicated that pendrin is expressed in ORNs and sustentacular cells in<br />

the epithelium, as well as in mitral cells in the olfactory bulb. In ORNs,<br />

sustentacular cells and mitral cells, pendrin is present in spherical,<br />

vesicle-like structures. The immunolabeling was specific since it was<br />

blocked with the pendrin peptide to which the antibody had been raised.<br />

Future studies are aimed at identifying the vesicles that express pendrin<br />

and in ascertaining the role of pendrin in regulating chloride<br />

concentration in the ORNs. This work was supported by NIH grant R01<br />

DC00926 (SJK) and DK 62809 (MS).<br />

73<br />

283 Poster [ ] Olfactory Sensory Neuron Physiology<br />

PLASMA MEMBRANE CALCIUM PUMPS IN THE MOUSE<br />

OLFACTORY AND VNO RECEPTOR CELLS<br />

Cusick M. 1, Chandran S. 1, Van Houten J. 1, Delay R. 2 1Biology,<br />

University of Vermont, Burlington, VT; 2Biology Department, University<br />

of Vermont, Burlington, VT<br />

Calcium (Ca) plays important roles in olfactory signaling and,<br />

there<strong>for</strong>e, it must be important to control intracellular Ca levels in<br />

olfactory and VNO sensory neurons by binding proteins, intracellular<br />

pumps, and extrusion from the cell by the Na+/Ca2+ transporter and<br />

plasma membrane Ca pumps (PMCAs). Mammals have 4 PMCA<br />

iso<strong>for</strong>ms, each with different kinetic properties to serve different<br />

physiological tasks. We are studying the distribution and roles of these<br />

pumps in olfactory and VNO receptor neurons using<br />

immunocytochemistry of epithelia and dissociated cells and calcium<br />

imaging. The pan-PMCA antibodies show distinct staining patterns in<br />

the olfactory and VNO epithelia: intense in the apical region of the<br />

olfactory epithelium but not in the respiratory epithelium; none apparent<br />

in the cilia; faint in the region of the neuron cell bodies. Use of specific<br />

antibodies shows: intense staining <strong>for</strong> iso<strong>for</strong>ms 1 and 2 in both VNO<br />

and olfactory epithelium; little or no staining <strong>for</strong> iso<strong>for</strong>ms 3 and 4 in<br />

either epithelium; weak to no staining of any iso<strong>for</strong>m in the respiratory<br />

epithelium. The most intense staining is in VNO <strong>for</strong> iso<strong>for</strong>ms 1 and 2<br />

and in olfactory epithelium.Dissociated olfactory sensory neurons<br />

(OSNs) show staining <strong>for</strong> iso<strong>for</strong>m 1 along the entire cell with intense<br />

staining where the dendrite meets the cell body. Staining <strong>for</strong> iso<strong>for</strong>m 2<br />

is prominent at the dendritic knob, cilia and cell body with less in the<br />

dendrite. Calcium imaging of isolated OSNs shows that calcium levels<br />

can return to basal levels after KCl or odorant stimulation even in the<br />

absence of transporter and SERCA pump function, which suggests a<br />

role <strong>for</strong> PMCAs in Ca removal in response to odor stimulation.<br />

Supported: NIH DC00721, P20RR16435, NCI PHS22435


284 Poster [ ] Olfactory Sensory Neuron Physiology<br />

OLFACTORY EPITHELIAL LOCALIZATION AND<br />

DENDRITIC MORPHOLOGY OF GOLF NEGATIVE<br />

OLFACTORY SENSORY NEURONS PROJECTING TO<br />

MEDIAL OLFACTORY BULB GLOMERULI IN THE LARVAL<br />

SEA LAMPREY (PETROMYZON MARINUS. L)<br />

Firby A.E. 1, Arbuckle W.J. 1, Zielinski B.S. 1 1Biological <strong>Sciences</strong>,<br />

University of Windsor, Windsor, Ontario, Canada<br />

Firby, A.E., Arbuckle, W.J., Zielinski, B.S.<br />

Abstract<br />

The objective of this study is to examine the dendritic morphology<br />

and olfactory epithelial distribution olfactory sensory neurons (OSNs)<br />

that project to spatially conserved medial glomeruli in the sea lamprey,<br />

an ancestral vertebrate with migratory and reproductive behavior<br />

mediated through pheromones (Li et al., 1995 J. Gen. Physiol. 105:<br />

567; Li et al., Science 2002 296: 138-141). These medial glomeruli<br />

lack Golf,expression; yet this GTP binding protein is localized in the<br />

remaining six glomerular territories (Frontini et al., J Comp Neurol<br />

465:27-37). Following micro-injection of fluorescent dextran amines<br />

into the medial glomeruli, dextran labeled OSNs were observed in the<br />

ventral hemisphere of the olfactory mucosa. The dendrite of these<br />

OSNs was long and slender, and the cell body was located in the basal<br />

half of the olfactory epithelium. In comparison, Golf- immunoreactive<br />

OSNs were widely distributed in the olfactory epithelium. The<br />

morphology of the Golf-immunoreactive OSNs included short, thick<br />

dendrites with cell bodies in the apical portion of the olfactory<br />

epithelium, and OSNs with slender dendrites and cell bodies in the basal<br />

half of the olfactory epithelium. This study shows that OSNs in the sea<br />

lamprey are dimorphic, and supports the idea that sub-populations of<br />

OSN terminals are distributed according to functional parameters in this<br />

species.<br />

Supported by: Great Lakes Fishery Commission, NSERC, University<br />

of Windsor Faculty of Graduate Studies<br />

74<br />

285 Poster [ ] Olfactory Sensory Neuron Physiology<br />

ELECTROPHYSIOLOGY OF SUSTENTACULAR CELLS IN<br />

MOUSE OLFACTORY EPITHELIUM (OE)<br />

Vogalis F. 1, Hegg C. 1, Lucero M. 1 1Physiology, University of Utah, Salt<br />

Lake City, UT<br />

Sustentacular cells (SCs) are exquisitely sensitive to P2Y receptor<br />

stimulation and generate robust increases in [Ca2+ ] implicating ATP as<br />

i<br />

an important paracrine regulator in the OE1 . Here we determined the<br />

electrical properties and responses to ATP of SCs in 250-µm OE slices<br />

from mice (P1-P4) using whole-cell techniques. Capacitances (C ) of m<br />

SCs averaged 17.9 ± 0.1 pF and input resistance (R ) was 173.1 ± 15.6<br />

in<br />

MΩ (n = 134). SCs generated a fast inward Na + current that was halfmaximally<br />

activated and inactivated at –52 mV and –86 mV<br />

respectively. Blocking gap junctions with 1 mM 1-octanol did not<br />

significantly alter C (20.9 ± 1.4 pF be<strong>for</strong>e vs. 20.7 ± 2.2 pF during 1-<br />

m<br />

octanol, n=11) while R increased significantly (p < 0.05, paired t-test)<br />

in<br />

from 140 ± 48 MΩ to 187 ± 58 MΩ. The poor electrical connectivity<br />

between SCs was confirmed by an absence of dye coupling in 8 of 10<br />

cells filled with Lucifer Yellow (0.2%). In the presence of 1-octanol,<br />

ATP (30-s, 20-50 µM) transiently decreased R by 10%, which then<br />

in<br />

increased by 20% during the 5 min wash. In 8 SCs, a 30-s application<br />

of ATP increased C by ~ 2.5% over 10 min. Our results indicate that<br />

m<br />

SCs generate inward Na + currents which, due to the low R of SCs,<br />

in<br />

cannot support action potentials. The low R is apparently not due to<br />

in<br />

coupling to neighboring cells. In addition, ATP elicits a biphasic<br />

response in the R of SCs and may trigger exocytosis as indicated by ~<br />

in<br />

400 fF increase in C . Further experiments will identify the resting<br />

m<br />

conductances of SCs and those elicited by ATP, as well as the Ca2+ -<br />

dependence of the ATP-mediated exocytosis. Funded by NIDCD<br />

DC02994toMTLand DC04953 toCCH.1.Heggetal.,J.Neurosci.23:<br />

8291-8301.


286 Poster [ ] Olfactory Sensory Neuron Physiology<br />

TRANSCRIPTS ENRICHED IN SENSORY NEURONS AND<br />

SUPPORTING CELLS OF THE OLFACTORY EPITHELIUM<br />

Yu T. 1, Mcintyre J.C. 1, Bose S.C. 1, Hardin D. 1, Mcclintock T.S. 1<br />

1Physiology, Cell and Molecular Sensory Systems Program, University<br />

of Kentucky, Lexington, KY<br />

Using a fluorescent LacZ substrate and flow cytometry, we collected<br />

LacZ+ and LacZ- cells from the olfactory epithelium of OMP-LacZ3<br />

mice (Walters et al., 1996, J Neurosci Res. 43:146). Olfactory marker<br />

protein was 100-fold enriched in the LacZ+ RNA, indicating successful<br />

purification of mature olfactory sensory neurons (OSNs).<br />

Representational difference analysis confirmed by quantitative RT-PCR<br />

identified 54 differentially distributed transcripts. The majority of these<br />

transcripts encode proteins that have no known function. In situ<br />

hybridization identified 11 transcripts expressed only by OSNs and<br />

vomeronasal sensory neurons (VSNs). Some were restricted to mature<br />

OSNs. Twelve additional transcripts were enriched in OSNs and<br />

VSNs, but were also expressed in other cells in the epithelium. One<br />

novel transcript was expressed in the OSNs, sustentacular cells, and<br />

Bowmans glands only within a restricted region of the epithelium. The<br />

six transcripts restricted to sustentacular cells and Bowman´s glands<br />

include several enzymes that metabolize odorants. We also identified a<br />

transcript, pancreatitis-associated protein, restricted to cells of the<br />

respiratory epithelium. As a group, these olfactory enriched transcripts<br />

provide useful markers of certain cell types within the epithelium and<br />

potential insights into the function of several of these cell types.<br />

Supported by R01 DC02736 and R01 AG18229.<br />

287 Poster [ ] Olfactory Sensory Neuron Physiology<br />

EXPRESSION PROFILING OF PHENOTYPICALLY<br />

IDENTIFIED OLFACTORY SENSORY NEURONS<br />

Mcintyre J.C. 1, Yu T. 1, Shetty R.S. 1, Sammeta N. 1, Smith M.A. 1,<br />

Mcclintock T.S. 1 1Department of Physiology, University of Kentucky,<br />

Lexington, KY<br />

GFP+ and GFP- cells were collected using flow cytometry from the<br />

olfactory epithelium of OMP-GFP mice (Potter et al., 2001, J Neurosci.<br />

21:24). RNA from these two populations of cells, and from the brain of<br />

OMP-GFP mice, was hybridized against Affymetrix MOE430 A and B<br />

GeneChip microarrays. Adenylyl Cyclase type III and G were<br />

&alphaolf<br />

present in the GFP+ RNA but not detected in the GFP- RNA, indicating<br />

that olfactory sensory neurons (OSNs) were enriched in the GFP+<br />

population. We identified 730 probe sets that were present only in the<br />

GFP+ population. Of these, 490 represent expressed sequence tags<br />

(ESTs), genes encoding proteins of unknown function, and unannotated<br />

genes. A wide range of functions were represented by the other 240<br />

sets. Transcripts were detected with roles in odorant detection (22), cell<br />

adhesion (10), metabolism and biosynthesis (21), transcription<br />

regulation (11), cell signaling (16), proliferation and differentiation<br />

(19), ion permeation and small molecule transport (24), and<br />

intracellular transport (9). Comparisons to previous experiments<br />

profiling transcripts expressed in the olfactory epithelium or in OSNs<br />

revealed overlaps with less than 5% of our list. Our data there<strong>for</strong>e<br />

appear to greatly expand the number of transcripts known to be<br />

enriched in OSNs.<br />

Supported by R01 DC02736 and R21 DC4507.<br />

75<br />

288 Poster [ ] Olfactory Sensory Neuron Physiology<br />

HOW SENSITIVE CAN A `BROADLY TUNED´ OLFACTORY<br />

RECEPTOR BE?<br />

Nickell T. 1 1Cell Biology, Neurobiology and Anatomy, University of<br />

Cincinnati, Cincinnati, OH<br />

There is an apparent paradox in the current literature of vertebrate<br />

olfaction. On the one hand, many kinds of evidence show that receptors<br />

are `broadly tuned´; each receptor responds to numerous compounds.<br />

On the other hand, mammals can detect very low concentrations of<br />

odorants and single receptor neurons may detect single odorant<br />

molecules. Broad tuning appears to require low-affinity binding of<br />

odorants; high sensitivity is most easily explained by high-affinity<br />

binding. Thus, there appears to be an inverse relation between<br />

sensitivity and `broad tuning´. We have used a simple model to explore<br />

this tradeoff.<br />

To obtain sensitivity near the theoretical limit of one molecule, one<br />

bound ligand--or one active receptor--must produce spikes in the<br />

olfactory neuron. Amplification mechanisms within the cell can<br />

accomplish this; however, in a receptor with low affinity and binding<br />

energy thermal fluctuations may produce a significant probability of the<br />

active state in the absence of a ligand. Unless the total spontaneous<br />

activity of all the receptors in the cell is well below threshold, these<br />

thermal fluctuations will produce spikes. This requirement determines a<br />

minimum energy difference between Active and Inactive states of the<br />

receptor and, hence, a minimum binding energy and affinity of the<br />

ligand (odorant).<br />

Calculations using plausible estimates of the number of olfactory<br />

receptors in a cell show that at a moderate binding energy, the<br />

probability of spontaneous receptor activity is low enough to permit a<br />

cellular response to a single active receptor.<br />

289 Poster [ ] Olfactory Sensory Neuron Physiology<br />

OLFACTORY `INTERFEROMETRY´ - NON-CONTIGUOUS<br />

DISTRIBUTIONS OF OLFACTORY RECEPTOR NEURONS<br />

EXPRESSING ONE OLFACTORY RECEPTOR<br />

Kauer J.S. 1, White J.E. 1 1Neuroscience, Tufts University School of<br />

Medicine, Boston, MA<br />

Following elucidation of olfactory receptor (OR) candidates (Buck<br />

and Axel, 1991), it was shown that olfactory receptor neurons (ORNs)<br />

expressing one OR (designated here ORNx´s) are positioned in a noncontiguous<br />

fashion within olfactory epithelial (OE) `zones´ (Ressler et<br />

al. 1993; Vassar et al.,1993). We have recently shown similar, noncontiguously<br />

positioned ORNx´s in salamander OE (Marchand et al.,<br />

2004, in revision). From a developmental perspective (how stem cells<br />

generate one ORNx population) and from a targeting perspective (how<br />

ORNx´s find glomerular targets in the bulb), it would seem more<br />

efficient that ORNx´s be in contiguous groups. The question thus arises<br />

as to what advantages might non-contiguous ORN placement af<strong>for</strong>d.<br />

In the present study, we hypothesize that non-contiguous ORNx<br />

placement provides the opportunity <strong>for</strong> odor-generated activity in each<br />

ORNx neuron to be compared by the receiving circuitry of the target<br />

glomerulus with respect to similarity. Responses from ORNx´s placed<br />

in different regions of the incoming air flow, should be similar because<br />

they arise in ORNs expressing the same OR. The degree to which such<br />

responses are seen as similar by glomerular circuits is a measure of<br />

signal robustness. Such comparisons, especially <strong>for</strong> near threshold<br />

signals, per<strong>for</strong>med in appropriate time windows (e.g. during sniffs), can<br />

provide <strong>for</strong> a kind of `interferometric´ process that could increase<br />

signal/noise ratio and improve the sensitivity (by averaging) of the<br />

ensemble, over individual ORN, responses.<br />

Supported by grants from the NIDCD and ONR.


290 Poster [ ] Olfactory Sensory Neuron Physiology<br />

ASSESSING AIRFLOW PARAMETERS IN RAT EOGS<br />

Scott J.W. 1, Acevedo H.P. 1 1Cell Biology, Emory University, Atlanta,<br />

GA<br />

The chromatographic model of olfactory response by Mozell and<br />

colleagues predicts that response size should vary with the interaction<br />

of flow rate and the physicochemical properties of odorants. We<br />

investigated this question in the rat nose by EOG recordings in the<br />

dorsal lateral recesses of the intact nasal cavity. Nasal flow rates of 50,<br />

100, 200, & 500 ml/min were used to characterize the response size and<br />

latency. We found that latency measurements in the lateral recess had to<br />

guard against complex wave<strong>for</strong>ms arising from stimulation at other<br />

sites. These wave<strong>for</strong>ms are artifactual because they do not appear if<br />

odorants are directly applied to the opened epithelium and they<br />

disappear if the dorsal recess is made unresponsive by prolonged<br />

exposure to low molecular weight esters. With isoamyl acetate<br />

stimulation, the response in the lateral recess grows more rapidly with<br />

increase in nasal flow rate than the response in the dorsal recess. The<br />

latency in the lateral recess also decreases as flow rate increases. This<br />

effect was present over at least a 3 log unit range of concentration,<br />

down to approximately 10-3 of vapor saturation. We suggest that the<br />

greater response latency in the lateral sites is indicative of slower<br />

airflow in the lateral recesses, which would favor odorant sorption<br />

be<strong>for</strong>e it reaches the recording site. As airflow increases, that loss of<br />

odorant due to sorption is decreased and the response increases<br />

markedly. This effect is smaller with odorants that normally evoke large<br />

lateral responses (hexane, α-terpinene, and limonene). We interpret<br />

these findings to mean that the sniff velocity may dynamically control<br />

both the timing and size of olfactory responses in different parts of the<br />

epithelium. Supported by NIH grants DC00113 & DC04710.<br />

291 Poster [ ] Olfactory Sensory Neuron Physiology<br />

ALTERED OLFACTORY SENSORY NEURON PHENOTYPE IN<br />

MUCOPOLYSACCHARIDOSES I AND VI<br />

Rawson N.E. 1, Wysocki L.M. 1, Dankulich L. 1, Gomez G. 2, Haskins M. 3<br />

1Monell Chemical Senses Center, Philadelphia, PA; 2Biology,<br />

University of Scranton, Scranton, PA; 3School of Veterinary Medicine,<br />

University of Pennsylvania, Philadelphia, PA<br />

Mucopolysaccharides (MPS) are thought to play a role in neural<br />

development and axon guidance. To assess their importance in the<br />

olfactory system, we studied the effects of two feline genetic models<br />

lacking different enzymes involved in MPS processing; one (MPSI) is<br />

associated with mental retardation, while the other (MPSVI) is not. We<br />

used tissue obtained at autopsy from unaffected control and affected<br />

cats. Stimulus-induced changes in intracellular calcium were studied<br />

using fluorescence imaging of live olfactory sensory neurons (OSNs).<br />

There were significantly fewer odorant responsive cells in the tissue<br />

from MPSI affected cats (n = 7; 1/57 ORNs), while about 25% of 149<br />

cells isolated from controls (n = 18) responded to a standard battery of<br />

13 odorants. In contrast, the frequency of responses to elevation of<br />

cAMP or membrane depolarization were similar in OSNs from control<br />

and affected cats. Fewer OSNs were obtained from animals with<br />

MPSVI (n=4; 11 OSNs), but they responded normally to odors,<br />

depolarization and cAMP elevation. These results indicate that the<br />

defect in alpha-L-iduronidase activity (MPSI), but not arylsulfatase B<br />

activity (MPSVI) alters the normal development or maintenance of the<br />

olfactory epithelium. These data suggest that abnormal MPS processing<br />

interferes with OSN function and that certain MPSs play a role in<br />

development or maintenance of the olfactory system. Funded in part by<br />

NIH DK25759 and RR02512 (MH).<br />

76<br />

292 Poster [ ] Olfactory Sensory Neuron Physiology<br />

BIOPHYSICAL MODEL OF OLFACTORY RECEPTOR<br />

NEURON (ORN) PAIRS REVEALS MECHANISM FOR GAP<br />

JUNCTION MEDIATED SYNCHRONIZED FIRING AT<br />

THRESHOLD ODOR CONCENTRATIONS<br />

Buntinas L. 1, Zhang C. 1, Restrepo D. 1 1Cell and Developmental<br />

Biology, University of Colorado Health <strong>Sciences</strong> Center, Denver, CO<br />

Recent studies in our laboratory have shown that multiple connexins<br />

are expressed in mature ORNs in mice, suggesting that gap junctional<br />

coupling might modify ORN function. A recent study in Necturus<br />

maculosus (Delay and Dionne Chem. Senses. 28:807, 2003) found only<br />

a small fraction of coupled ORNs arguing against a general functional<br />

role <strong>for</strong> gap junctions. However, sparse coupling does not imply lack of<br />

a functional role of gap junctions in sub-groups of neurons expressing<br />

the same olfactory receptor. If connexins are only expressed in ORNs<br />

expressing certain olfactory receptors, the function of those ORNs<br />

could be modulated by gap junctional coupling. We <strong>for</strong>mulated a<br />

Hodgkin and Huxley compartmental model of two gap junction coupled<br />

ORNs in order to create the theoretical framework necessary to test the<br />

hypothesis that gap junctions modulate function of specific subgroups<br />

of ORNs. We asked how the strength of the coupling would change the<br />

odor-induced activity of the coupled neurons. We report that at<br />

intermediate coupling strength, ORNs of different odor specificity<br />

display synchronized firing at threshold odor concentrations. At higher<br />

concentrations, the firing is not synchronized and differs substantially<br />

between the cell that is stimulated with the cognate odor and the<br />

adjacent cell. This mechanism would allow <strong>for</strong> increased odor<br />

sensitivity of certain odors without compromising odor quality<br />

discrimination.<br />

This work was supported by NIH grants DC00566, DC04657 (DR),<br />

DC04952 (CZ) and DC006542 (LB).<br />

293 Poster [ ] Sweet Taste<br />

REDUCTION OF SWEET-SUPPRESSING EFFECTS OF<br />

GURMARIN BY KALLIKREINS INCREASED IN THE<br />

SUBMANDIBULAR SALIVA OF RATS FED GYMNEMA-<br />

CONTAINING DIET<br />

Yamada A. 1, Katsukawa H. 2, Sugita D. 3, Ninomiya Y. 1 1Kyushu<br />

University, Fukuoka, Japan; 2Physiology, Asahi University, Gifu,<br />

Japan; 3Central Laboratory, Lotte Co., LTD., Saitama, Japan<br />

A tropical plant, gymnema sylvestre (gymnema), contains gurmarin<br />

that selectively inhibits responses to sweet substances in rodents. The<br />

present study investigated possible interaction between gurmarin and<br />

the submandibular saliva in rats fed diet containing gymnema. At 1-2<br />

days after the start of the gymnema diet, preference <strong>for</strong> saccharin and<br />

D-phenylalanine decreased and subsequently returned closely to the<br />

control level within several days. Electrophoretic analyses<br />

demonstrated that relative amounts of two proteins in the saliva clearly<br />

increased in rats fed the gymnema diet. Rats previously given section of<br />

the bilateral glossopharyngeal nerve, however, showed no such salivary<br />

protein induction, suggesting importance of sensory in<strong>for</strong>mation <strong>for</strong> the<br />

protein induction. Analyses of amino acid sequence indicate that two<br />

proteins are rat kallikrein 2 (rK2) and rat kallikrein 9 (rK9). rK2 and<br />

rK9, a family of serine proteases, have resemble cleavage sites in the<br />

protein substrates of which comparable residue is also contained in<br />

sequence of gurmarin. Finally, kallikreins purified from saliva clearly<br />

inhibited the immunoreaction between gurmarin and antigumarin<br />

antiserum. These results suggest that rK2 and rK9 increased by the<br />

gymnema diet via oral sensory system cleave gurmarin and reduce its<br />

sweet suppressing effect in rats.


294 Poster [ ] Sweet Taste<br />

BEHAVIORAL TESTING OF THE INTERACTION OF SWEET<br />

TASTE AND SOLUTION TEMPERATURE IN THE RAT<br />

Denbleyker M. 1, Taylor P.A. 1, Smith P. 2, Smith J.C. 1 1Psychology,<br />

Florida State University, Tallahassee, FL; 2Psychology, Florida<br />

Southern University, Lakeland, FL<br />

Electrophysiological recordings have shown that there is an<br />

interaction between taste and temperature in the rat, but there is scant<br />

behavioral evidence to support this finding. The present study was<br />

conducted to test the licking behavior of non-deprived rats in a shortterm<br />

test using different concentrations of nutritive and non-nutritive<br />

sweeteners across temperatures ranging from 10° to 40°C. A testing<br />

apparatus was developed to control the temperature of a solution with a<br />

resolution of 1° C. An infrared beam was passed across the opening to<br />

the sipper tube, so when the rat's tongue broke the beam, licks could be<br />

counted. Shutters could be opened, allowing access to the sipper tubes.<br />

During daily testing sessions, eight 30s taste/temperature trials were<br />

given by opening the shutter, allowing the rat access to a sucrose or<br />

saccharin solution. The inter-trial-interval was 90s, allowing time <strong>for</strong><br />

the temperature of the solution to be raised 5° <strong>for</strong> the subsequent trial.<br />

On a given day, the concentration of the sucrose or saccharin was held<br />

constant and the temperature was varied from 10° to 40° in an<br />

ascending manner across the 8 trials. Concentrations used <strong>for</strong> sucrose<br />

were 0.00075 to 0.25 M and <strong>for</strong> saccharin 0.001 to 0.066 M. The<br />

software analysis program allowed <strong>for</strong> a microanalysis of the licking<br />

behavior during the 30s presentations. As has been shown, licking<br />

increases as a function of concentration <strong>for</strong> sucrose and is expressed as<br />

an inverted U-function <strong>for</strong> saccharin. In all cases, licking decreases as<br />

the temperature of the solution is increased above 22°C.<br />

295 Poster [ ] Sweet Taste<br />

MOLECULAR STUDIES OF SWEET TASTE RECEPTOR<br />

FUNCTION<br />

Jiang P. 1, Liu Z. 2, Benard L. 2, Snyder L. 1, Margolskee R. 2, Max M. 1<br />

1Department of Physiology and Biophysics, Mount Sinai School of<br />

Medicine, New York, NY; 2Department of Physiology and Biophysics,<br />

Mount Sinai School of Medicine, Howard Hughes Medical Institute,<br />

New York, NY<br />

Heterologous expression of T1R2 plus T1R3 yields a functional<br />

“sweet receptor” that is responsive to a diverse range of sweet tasting<br />

ligands. Sweet tastants include sugars (e.g. glucose, fructose, sucrose),<br />

sugar alcohols, small molecule artificial sweeteners (e.g. saccharin and<br />

acesulfame K) and certain proteins (e.g. brazzein, monellin and<br />

thaumatin). There is no common structure that unites all of these<br />

diverse compounds, although several attempts have been made to infer<br />

a universal sweet motif. Brazzein, like monellin and thaumatin, is a<br />

naturally occurring protein that humans, apes and old world monkeys<br />

perceive as intensely sweet, but which is not preferred by, and is<br />

apparently tasteless to, other species such as new world monkeys, rats<br />

and mice.<br />

Using mouse-human T1R chimeras, site-directed mutagenesis and<br />

calcium imaging of heterologously-expressed T1R2 + T1R3 we have<br />

determined the molecular basis <strong>for</strong> this species-specific sensitivity: we<br />

have identified a site within human TlR3 that is required <strong>for</strong> brazzein to<br />

stimulate T1R2 + T1R3. Other mutations in this same region of human<br />

T1R3 had effects on receptor activity toward monellin, and in some<br />

cases, overall receptor responsivity toward most or all sweet<br />

compounds. This suggests that this region of T1R3 may play a role in<br />

both ligand binding (especially <strong>for</strong> the protein sweeteners) and in the<br />

transition between inactive and active states of the receptor.<br />

77<br />

296 Poster [ ] Sweet Taste<br />

ALLELIC VARIATION OF THE TAS1R3 TASTE RECEPTOR<br />

GENE SELECTIVELY AFFECTS BEHAVIORAL AND<br />

NEURAL TASTE RESPONSES TO SWEETENERS IN THE F2<br />

HYBRIDS BETWEEN C57BL/6BYJ AND 129P3/J MICE<br />

Inoue M. 1, Reed D.R. 2, Li X. 2, Tordoff M.G. 2, Bachmanov A.A. 2,<br />

Beauchamp G.K. 2 1Tokyo University of Pharmacy and Life Science,<br />

Tokyo, Japan; 2Monell Chemical Senses Center, Philadelphia, PA<br />

Recent studies have shown that the T1R3 receptor protein encoded<br />

by the Tas1r3 gene is involved in transduction of sweet taste. To assess<br />

ligand specificity of the T1R3 receptor, we analyzed the association of<br />

Tas1r3 allelic variants with taste responses in mice. In the F2 hybrids<br />

between the C57BL/6ByJ (B6) and 129P3/J (129) inbred mouse strains,<br />

we determined genotypes of markers on chromosome 4, where Tas1r3<br />

resides, measured consumption of taste solutions presented in the twobottle<br />

preference tests, and recorded integrated responses of the chorda<br />

tympani gustatory nerve to lingual application of taste stimuli. For<br />

intakes and preferences, significant linkages to Tas1r3 were found <strong>for</strong><br />

the sweeteners sucrose, saccharin and D-phenylalanine, but not glycine.<br />

For chorda tympani responses, significant linkages to Tas1r3 were<br />

found <strong>for</strong> the sweeteners sucrose, saccharin, D-phenylalanine, Dtryptophan<br />

and SC-45647, but not glycine, L-proline, L-alanine or Lglutamine.<br />

No linkages to distal chromosome 4 were detected <strong>for</strong><br />

behavioral or neural responses to non-sweet quinine, citric acid, HCl,<br />

NaCl, KCl, monosodium glutamate (MSG), inosine 5´-monophosphate<br />

(IMP) or ammonium glutamate. These results demonstrate that allelic<br />

variation of the Tas1r3 gene affects gustatory neural and behavioral<br />

responses to some but not all sweeteners. This study describes the range<br />

of ligand sensitivity of the T1R3 receptor using an in vivo approach.<br />

Supported by NIH grant DC00882 (GKB).<br />

297 Poster [ ] Taste Psychophysics<br />

PERCEPTUAL VARIANCE: HOW DISCRIMINATION<br />

METHODS BECOME LESS DISCRIMINATING<br />

Lee H. 1, Jeon S. 2, Kim K. 3, O'Mahony M. 4 1Food Science &<br />

Technology, University of Cali<strong>for</strong>nia, Davis, Davis, CA; 2Dept Food &<br />

Nutritional Science, Ewha Womans University, Seoul, South Korea;<br />

3Dept. Food & Nutritional Science, Ewha Womans University, Seoul,<br />

South Korea; 4Food Science and Technology, University of Cali<strong>for</strong>nia,<br />

Davis, Davis, CA<br />

Discrimination tests are incorporated into threshold measurement and<br />

are used as difference tests, in the sensory evaluation of food. Yet, the<br />

various methods are not equivalent. This can be understood in terms of<br />

perceptual distributions (signal & noise) and d'. For a given perceptual<br />

difference, the greater the variance of these distributions, the smaller<br />

will be the value of d'. There are several sources of perceptual variance<br />

that can affect discrimination: adaptation and the sequence of tasting,<br />

memory and <strong>for</strong>getting, and criterion variation. The goal of this study<br />

was to investigate these effects, using discrimination between purified<br />

water and `threshold´ concentrations of NaCl. Firstly, the 2-AFC,<br />

triangle, duo-trio, and same-different methods were compared. Values<br />

of d' were computed and used <strong>for</strong> statistical analysis. After warm-up,<br />

which had the effect of stabilizing its criterion, memory effects<br />

rendered the same-different test as sensitive as the 2-AFC, despite its<br />

less advantageous sequence effects. Both were significantly more<br />

sensitive than the duo-trio or triangle methods. However, without<br />

warm-up criterion variation reduced the sensitivity of the samedifferent<br />

method to that of the triangle. Secondly, the effects of<br />

adaptation were further demonstrated using the two possible 3-AFC<br />

tests (NaCl-odd vs. water-odd) and varying their relative sensitivity by<br />

manipulating interstimulus rinses. Either 3-AFC elicited significantly<br />

higher d' values when the interstimulus rinses were different from the<br />

stimuli than when they were the same.<br />

Funding: UC Davis, Ewha WU


298 Poster [ ] Taste Psychophysics<br />

BIMODAL DISTRIBUTION OF SUCROSE OCTAACETATE<br />

(SOA) BITTER TASTE SENSITIVITY, AND HERITABILITY<br />

OF THIS TRAIT AMONG TWINS<br />

Tharp A.A. 1, Alarcon S.M. 1, Tharp C.D. 1, Reed D.R. 1, Breslin P.A. 1<br />

1Monell Chemical Senses Center, Philadelphia, PA<br />

Sucrose octaacetate (SOA) tastes bitter to most tasters but the degree<br />

of perceived bitterness varies considerably. SOA is one of the most<br />

extensively investigated bitter taste traits in mice and has led to the<br />

identification of the Soa locus on mouse chromosome 6. To determine<br />

whether human sensitivity to the bitterness of SOA is also a genetic<br />

trait, we screened 130 mono- and dizygotic twins <strong>for</strong> their bitterness<br />

recognition threshold using a modified Harris-Kalmus sort test. In<br />

addition, these same subjects tasted and rated the bitterness intensity of<br />

six concentrations (including water) of SOA twice on the general<br />

Labeled Magnitude Scale. Monozygotic twins share 100% of their<br />

alleles in common whereas dizygotic twins share 50% of their genes on<br />

average. Thus, if a perceptual/behavioral trait has a significant genetic<br />

component, then monozygotic twins should be more similar to each<br />

other <strong>for</strong> this trait than are dizygotic twins. We found that the perceived<br />

bitterness of SOA was more similar among monozygotic twins than<br />

dizygotic twins <strong>for</strong> both the Harris-Kalmus test and the bitterness<br />

ratings. We further found that bitterness ratings were reliable and, at<br />

low concentrations, <strong>for</strong>med a bimodally distributed frequency<br />

histogram due to a subset of the population that cannot taste SOA until<br />

higher concentrations. At higher SOA intensities, the distribution<br />

appears more unimodally distributed. We also determined that the<br />

intensity ratings at low concentrations and the Harris-Kalmus bitterness<br />

recognition threshold levels were correlated, which validates these<br />

observations. Human Chr 12 syntenic to the Soa locus on mChr 6<br />

should yield candidate genes. Reasearch funded by DC02995 to PASB.<br />

299 Poster [ ] Taste Psychophysics<br />

GUSTATORY RESPONSE TIMES TO INTENSITY AND<br />

HEDONIC JUDGMENTS<br />

Veldhuizen M.G. 1, Kroeze J.H. 2 1Taste and Smell Laboratory,<br />

University of Utrecht, Utrecht, Netherlands; 2Taste and Smell<br />

Laboratory, University of Utrecht, Utrecht, Netherlands<br />

Choice response times of intensity and hedonic judgments were<br />

compared. Subjects made <strong>for</strong>ced choice paired comparisons of orange<br />

lemonades with various concentrations of added quinine sulfate.<br />

Subjects were instructed to focus either on intensity or on pleasantness.<br />

Computerized administration of the stimuli and registration of the<br />

responses enabled response times measurements. Gustatory response<br />

times to intensity and hedonic judgments were compared in a withinsubjects<br />

ANOVA. Preliminary results of 24 subjects indicate that a<br />

focus on intensity or hedonic aspects had no effect on response times<br />

(F[1,23] = 2.451, p = .131), whereas there was an effect of<br />

concentration on response times (F[3.69] = 6.164, p = .001). The<br />

similarity of the response times <strong>for</strong> intensity and pleasantness may be<br />

due to the simultaneous availability of hedonic and intensity<br />

in<strong>for</strong>mation to cognitive processing. Such a conclusion may be more<br />

compatible with a model of parallel processing of hedonic and<br />

perceptual in<strong>for</strong>mation than with serial models of hedonic and<br />

perceptual processing.<br />

Funding by Helmholtz Institute, University of Utrecht<br />

78<br />

300 Poster [ ] Taste Psychophysics<br />

THERMAL TASTE IS ASSOCIATED WITH GENERALLY<br />

HIGHER TASTE RESPONSIVENESS.<br />

Green B. 1, George P. 2 1The John B. Pierce Laboratory and Yale School<br />

of Medicine, New Haven, CT; 2The John B. Pierce Laboratory, New<br />

Haven, CT<br />

It was recently found that <strong>for</strong> some people, temperature alone can<br />

stimulate taste sensations (Thermal taste). To investigate a possible<br />

source of individual differences in thermal taste we tested whether<br />

sucrose sweetness, which can be modulated by temperature, would be<br />

rated higher by thermal tasters (n=14) than by thermal non-tasters<br />

(n=14). Ss used the gLMS to rate taste intensity on the tongue tip <strong>for</strong><br />

three concentrations of sucrose, with saccharin and NaCl included as<br />

controls. Contrary to the hypothesis, thermal tasters gave higher ratings<br />

to all three stimuli [F(1,26)=21.0, p


302 Poster [ ] Taste Psychophysics<br />

SUCROSE AND SODIUM CHLORIDE SELF-ADAPTATION<br />

USING “TASTE”<br />

Ashkenazi A. 1, Gent J.F. 2, Marks L.E. 1, Frank M.E. 3 1J. B. Pierce<br />

Laboratory and Yale University, New Haven, CT; 2Yale University,<br />

New Haven, CT; 3University of Connecticut, Farmington, CT<br />

Recently we developed a computer-controlled, automated, open flow<br />

system <strong>for</strong> gustatory research (Ashkenazi, Fritz, Buckley, and Marks, in<br />

press). Using pressurized air to control delivery of the solutions, this<br />

system provides precise temporal control over as many as 16 possible<br />

stimuli. In the current study we tested whether a common finding in the<br />

taste domain, that of self-adaptation, can be replicated using TASTE<br />

(Temporal Automated System <strong>for</strong> Taste Experiments). Subjects rated<br />

the intensity of either 0.5 M sucrose or 0.5 M sodium chloride on a<br />

computerized version of the Labeled Magnitude Scale (Bartoshuk,<br />

Duffy, Fast, Green, Prutkin, and Snyder, 2003). Intensity ratings were<br />

greater <strong>for</strong> both test stimuli when preceded by a 30-sec water rinse<br />

compared to a 30-sec self-adaptation rinse. Further, the pattern of<br />

results was similar when effects of order were controlled. These<br />

findings, similar to findings obtained with traditional methods of<br />

stimulus delivery, support the operational validity of TASTE <strong>for</strong> studies<br />

of adaptation. This work was supported by NIH grants R01 DC004849-<br />

01 to MEF and R01 DC00271-16 to LEM<br />

303 Poster [ ] Taste Psychophysics<br />

PERCEIVED INTENSITY FUNCTIONS GENERATED UNDER<br />

SIMULATED FMRI SCANNING CONDITIONS<br />

Haase L.B. 1, Cerf-Ducastel B. 1, Mellinger C. 1, Jacobson A. 1, Murphy<br />

C. 2 1Psychology, San Diego State University, San Diego, CA;<br />

2Psychology, San Diego State University, University Of Cali<strong>for</strong>nia, San<br />

Diego, San DIego, CA<br />

Currently, research on the psychophysics of taste has become<br />

increasingly important; e.g., <strong>for</strong> interpretation of functional MRI data.<br />

However, an important question is how slopes of functions relating<br />

perceived intensity to concentration of taste stimuli measured under<br />

conditions outside of the scanning environment will be related to slopes<br />

of intensity functions <strong>for</strong> stimuli presented under the severe constraints<br />

on stimulus delivery that apply in fMRI experiments. The current study<br />

investigated perceived intensity across a number of taste stimuli at<br />

various concentrations, in healthy young-adults, with a stimulation<br />

protocol adapted <strong>for</strong> fMRI scanning. Stimuli were presented at regular<br />

intervals in small boluses (.03ml) to minimize swallowing when the<br />

subject is lying on the back inside the scanner. Intensity was measured<br />

using Green´s Labeled Magnitude Scale. Slopes of intensity functions<br />

<strong>for</strong> stimuli presented in the simulated scanning conditions of the present<br />

study were compared to slopes from previous studies using the dorsal<br />

flow and "sip and spit" techniques. Results indicate that the slopes of<br />

intensity functions <strong>for</strong> Sucrose, NaCl, Caffeine, Saccharin, and MSG in<br />

the simulated scanning conditions were similar to the values <strong>for</strong> slopes<br />

from dorsal flow studies and generally lower than in studies done with<br />

the sip and spit method. Previous psychophysical taste research using<br />

the dorsal flow technique may be particularly useful <strong>for</strong> interpretation<br />

of future fMRI experiments.<br />

Supported by NIH Grant # AG04085 to Claire Murphy.<br />

79<br />

304 Poster [ ] Taste Psychophysics<br />

SUPERTASTING IS NOT EXPLAINED BY THE PTC/PROP<br />

GENE<br />

Bartoshuk L.M. 1, Davidson A. 2, Kidd J. 2, Kidd K.K. 2, Speed W. 3,<br />

Pakstis A. 2, Reed D. 4, Snyder D. 1, Duffy V.B. 5 1Surgery, Yale<br />

University, New Haven, CT; 2Genetics, Yale University School of<br />

Medicine, New Haven, CT; 3Genetics, Yale University, New Haven, CT;<br />

4Monell Chemical Senses Center, Philadelphia, PA; 5Dietetics<br />

Program, University of Connecticut, Storrs, CT<br />

An important gene (TAS2R38) contributes to PTC/PROP perception<br />

(Kim et al, 2003). Two common <strong>for</strong>ms of this putative taste receptor are<br />

defined by single nucleotide polymorphisms that result in three amino<br />

acid substitutions: Pro49Ala, Ala262Val, and Val296Ile, leading to<br />

haplotypes PAV and AVI. The ancestral human haplotype is Proline-<br />

Alanine-Valine (PAV) which was associated with lower PTC<br />

thresholds (tasters); AVI (I=Isoleucine) was associated with higher PTC<br />

thresholds (nontasters). Subjects provided PROP thresholds (N=54),<br />

rated the bitterness of PROP (0.032-3.2 mM) with the general Labeled<br />

Magnitude Scale and were genotyped <strong>for</strong> the PTC/PROP gene (N=91).<br />

Threshold data corroborated those presented by Kim et al; PAV/PAV<br />

thresholds were lowest, AVI/AVI thresholds highest and PAV/AVI<br />

thresholds intermediate. However, suprathreshold functions showed<br />

that PAV/PAV tasters perceived significantly but only slightly more<br />

bitterness than did PAV/AVI tasters; AVI/AVI nontasters perceived the<br />

least bitterness. Of critical note, PAV/PAV tasters are not identical with<br />

supertasters defined by psychophysical criteria. Some of the factors (in<br />

addition to the PTC/PROP gene) likely to contribute to the perceived<br />

bitterness of PROP are density of fungi<strong>for</strong>m papillae, other bitter genes<br />

and pathology. The distinction between medium and supertasters plays<br />

an important role in health outcomes related to oral perception.<br />

(NRICGP/USDA 2002-00788, DC00283, GM 57672).<br />

305 Poster [ ] Taste Psychophysics<br />

CHILDHOOD TOBACCO EXPOSURE INCREASES OBESITY<br />

RISK IN ADULT MEN<br />

Snyder D.J. 1, O'Malley S.S. 2, Mckee S. 2, Bartoshuk L.M. 1 1Surgery,<br />

Yale University, New Haven, CT; 2Psychiatry, Yale University, New<br />

Haven, CT<br />

Maternal smoking during pregnancy promotes childhood obesity, but<br />

its impact on adult body mass remains unclear. Notably, these findings<br />

fail to consider the role of tobacco use in the home, where chronic<br />

exposure throughout childhood may confer long-term health risk. Early<br />

tobacco exposure promotes childhood ear infections, which alter oral<br />

sensation by damaging the chorda tympani; <strong>for</strong> male supertasters of<br />

PROP (6-n-propylthiouracil), such changes may encourage fat intake<br />

and adult-onset obesity. Consistent with this model, we show that<br />

postnatal tobacco exposure correlates with increased body mass index<br />

(BMI) in adult men. Participants in a smoking cessation program<br />

(N=288) provided height, weight, family smoking history, and patterns<br />

of tobacco and alcohol use and abstinence. Tobacco exposure produced<br />

significant differences in men only; women were unaffected. Adult men<br />

raised in homes with 2+ smokers during ages 1-10 had higher BMIs<br />

than did men raised among fewer smokers; they were also more likely<br />

to be overweight. Although BMI often rises with smoking cessation,<br />

these men were willing to tolerate extreme gains leading to clinical<br />

obesity. Childhood exposure to maternal smoking contributed to higher<br />

rates of quit-related BMI gain, but prenatal exposure was unrelated to<br />

any adult BMI measure. These data suggest that postnatal tobacco<br />

exposure advances obesity risk in adult men by supporting pathologic<br />

changes in oral sensation. (Supported by the Transdisciplinary Tobacco<br />

Use Research Center at Yale and grants from NIDA/NCI (SSO), and<br />

NIDCD (LMB).)


306 Poster [ ] Taste Psychophysics<br />

THE INFLUENCE OF HEAD TRAUMA (HT), OTITIS MEDIA<br />

(OM), AND TONSILLECTOMY ON ORAL SENSATION, FAT<br />

ACCEPTANCE, AND BODY MASS INDEX (BMI)<br />

Chapo A. 1, Alex J. 2, Coelho D. 1, Duffy V.B. 3, Snyder D. 1, Bartoshuk L. 1<br />

1Surgery, Yale University, New Haven, CT; 2Otolaryngology, Lahey<br />

Clinic Medical Center, Burlington, MA; 3Dietetics, University of<br />

Connecticut, Storrs, CT<br />

Due to the close proximity of cranial nerve IX to the tonsillar bed,<br />

taste could be damaged by tonsillectomy. The present study explores<br />

oral sensation, food preferences, and BMI related to tonsillectomy and<br />

other taste-related pathology. <strong>Lecture</strong> attendees (n=674) reported<br />

weight, height, and pathology (OM, HT, and tonsillectomy) and rated<br />

remembered sensations, butterscotch candy sweetness, and acceptance<br />

of 26 foods on the general Labeled Magnitude Scale (gLMS). A<br />

statistically reliable group of 9 high-fat foods was positively associated<br />

with BMI. Females aged 25-65 who demonstrated proper use of the<br />

gLMS (most extreme light and sound greater than/equal to strong) were<br />

included and classified as follows: controls reported no pathology,<br />

“OM/HT” reported OM and/or HT (indicating VII loss), “TONs”<br />

reported tonsillectomies (indicating IX loss), and “COMBOs” reported<br />

OM and/or HT with tonsillectomy (indicating VII and IX loss).<br />

Compared to controls and TONs, COMBOs perceived less candy<br />

sweetness, suggesting taste loss; in those aged 40+, COMBOs liked fat<br />

foods more and had greater BMIs. OM/HTs perceived less sweetness<br />

than controls. In those aged 40+, OM/HTs tended to like fat foods more<br />

than controls but the two groups showed no differences in BMI.<br />

Summary: Damage to VII and IX may release inhibition on oral<br />

somatosensation; subsequent changes in fat acceptance may modulate<br />

obesity risk. Damage to VII alone may produce intermediate effects.<br />

(DC00283)<br />

307 Slide [ ] Taste Psychophysics<br />

GUSTATORY RESPONSES TO UNILATERAL<br />

GLOSSOPHARYNGEAL NERVE DAMAGE<br />

Pitovski D.Z. 1, Goins M. 1 1WFU Smell and Taste Center, Department<br />

of Otolaryngology, Wake Forest University, Winston-Salem, NC<br />

At our Wake Forest University Smell and Taste Center we have been<br />

referred several patients (in a six-month period) with the complaint of<br />

taste distortion following unilateral tonsillectomy. We report a patient<br />

that complains of taste distortion following a right tonsillectomy <strong>for</strong><br />

unilateral tonsillar hypertrophy. After a complete clinical evaluation<br />

and taste testing it was found that the patient suffered an injury to the<br />

right lingual branch of the glossopharyngeal nerve. The close<br />

anatomical relationship between the palatine tonsil and lingual branch<br />

of the glossopharyngeal nerve makes the nerve vulnerable during<br />

tonsillectomy. This injury has caused the patient to suffer ageusia to the<br />

right posterior one-third of the tongue compensated by a contralateral<br />

phantogeusia (phantom taste) with clinical dysgeusia. The phantogeusia<br />

was abolished by application of anesthetic to the area where the<br />

phantom was perceived. We propose that the phantogeusia is the result<br />

of release-of-inhibition in the contralateral glossopharyngeal nerve.<br />

Taste distortion (including, phantogeusia and dysgeusia) after<br />

tonsillectomy is rarely reported as complication, but has a significant<br />

impact on quality of life. This preliminary study examines the taste<br />

distortion presence as a complication following glossopharyngeal nerve<br />

damage during tonsillectomy.<br />

80<br />

308 Slide [ ] Cortical Signal Processing<br />

FLEXIBILITY, NOT CONTENT, OF CUE REPRESENTATIONS<br />

IN ABL DEPENDS ON INPUT FROM OFC<br />

Schoenbaum G. 1, Saddoris M.P. 2, Gallagher M. 2 1Department of<br />

Anatomy and Neurobiology, University of Maryland at Baltimore,<br />

Baltimore, MD; 2Department of Psychological and Brain <strong>Sciences</strong>,<br />

Johns Hopkins University, Baltimore, MD<br />

Interactions between orbitofrontal cortex (OFC) and basolateral<br />

amygdala (ABL) are critical to the encoding and use of in<strong>for</strong>mation<br />

regarding incentive value. We have reported that neurons in ABL<br />

develop differential firing to odor cues paired with appetive and<br />

aversive tastant outcomes. These differential responses reflect the<br />

acquired value of the cue and the predicted outcome. To assess the<br />

influence of input from OFC, we recorded from ABL neurons in rats<br />

with ipsilateral neurotoxic lesions of OFC as the rats were learning and<br />

reversing novel 2-odor discrimination problems. Correlates of these<br />

neurons were compared with those of ABL neurons recorded in rats<br />

with sham lesions. We compared activity on S+ and S- trials during<br />

odor sampling, responding and outcome presentation. We found that<br />

OFC lesions had little effect on the proportions of cue-selective neurons<br />

observed or on how selectivity emerged in these neurons with learning.<br />

Nor did the OFC lesions affect the proportion of the cue-selective<br />

neurons that also exhibited differential activity be<strong>for</strong>e and during<br />

outcome presentation. OFC lesions did, however cause a dramatic<br />

reduction in the flexibility of these representations after reversal. While<br />

53% of the cue-selective neurons reversed odor preference after<br />

reversal in intact rats, only 15% did so in rats with OFC lesions. These<br />

results suggest that although input from OFC may not be fundamental<br />

to the ability of ABL to represent cue value, this input does appear to be<br />

critical <strong>for</strong> the rapid flexibility of those representations, consistent with<br />

the deficit in reversal learning that is the hallmark of OFC lesions in<br />

rats and primates. Supported by K08-AG00882 and R01-DA015718<br />

(GS) and R01-MH60179 (MG).


309 Slide [ ] Cortical Signal Processing<br />

BRAIN ACTIVATION PATTERN IN RESPONSE TO<br />

OLFACTORY RECOGNITION MEMORY<br />

Cerf-Ducastel B. 1, Chen M. 2, Abou E. 3, Haas L. 3, Murphy C. 1<br />

1Psychology, San Diego State University, San Diego, CA; 2University of<br />

Cali<strong>for</strong>nia, San Diego, san diego, CA; 3San Diego State University, san<br />

Diego, CA<br />

Ten young subjects (age 20 to 25 y) participated in a functional<br />

Magnetic Resonance Imaging (fMRI) study with a 3T MR scanner.<br />

Be<strong>for</strong>e entering the scanner subjects were presented with 16 familiar<br />

odors. They then were scanned <strong>for</strong> 3 runs during which they were<br />

presented with words on a screen every 4 sec which were either names<br />

of odors previously presented (targets) or names of new odors (foils).<br />

Subjects responded by pressing either button 1 if they recognized the<br />

odor or button 2 if not. Each run alternated 4 `ON´ periods containing 7<br />

targets and 2 foils (36 s) and 4 `OFF´ periods with 7 foils and 2 targets<br />

(36 s). Data were processed with AFNI (Cox 1996) and compared ON<br />

and OFF periods, extracting regions that responded to cross-modal<br />

olfactory recognition memory. Group analysis showed that during the<br />

first run activated regions included right hippocampus,<br />

piri<strong>for</strong>m/amygdalar area, superior temporal gyrus, anterior cingulate<br />

gyrus, inferior frontal/orbito frontal gyrus, superior/medial frontal<br />

gyrus, and bilateral parahippocampal gyrus, inferior parietal lobule,<br />

supramarginal gyrus, cerebellum, lingual/fusi<strong>for</strong>m area, and<br />

middle/posterior cingulate gyrus. Region of interest analysis showed<br />

that degree of activation significantly decreased from run 1 to run 3 in<br />

the right hippocampus, fusi<strong>for</strong>m and lingual gyrus, parahippocampal<br />

gyrus and middle frontal gyrus but not in other regions, suggesting that<br />

these regions sustain a specific function in olfactory recognition<br />

memory that attenuates as foils become more familiar with repeated<br />

presentation. Supported by NIH grant AG04085 to CM.<br />

310 Slide [ ] Cortical Signal Processing<br />

CONTEXT DEPENDANT ACTIVITY IN PRIMARY<br />

OLFACTORY CORTEX OF HUMANS<br />

Sobel N. 1, Zelano C. 1, Mainland J. 1, Porter J.A. 1, Johnson B. 1, Bremner<br />

E. 1, Bensafi M. 1, Khan R. 1 1Neuroscience, University of Cali<strong>for</strong>nia,<br />

Berkeley, Berkeley, CA<br />

Attentional modulation of neural activity patterns has been<br />

demonstrated in primary visual and auditory cortex. Similarly, activity<br />

in the rat olfactory bulb reflects non-odor cues that are merely<br />

associated with odor content, and activity in rat olfactory cortex<br />

reflects the motivational significance of an odor. To ask whether<br />

contextual modulation is evident in activity patterns of human primary<br />

olfactory cortex we used functional magnetic resonance imaging to<br />

measure sniff-induced neural activity in and out-of an olfactory context.<br />

Each trial of an event related design began with an auditory primer <strong>for</strong><br />

"task detection" or "task inhalation". In "task detection" subjects took<br />

one sniff and determined whether an odorant was present or not. Odor<br />

was present on half of these trials. In "task inhalation" subjects also<br />

took one sniff, but knew in advance that an odor would never be<br />

present. Thus, the only difference between sniffs in "task inhalation",<br />

and the no-odor sniffs of "task detection" was that in the latter condition<br />

subjects were exploring <strong>for</strong> the presence of odor. Real-time<br />

measurement of nasal respiration assured sniffs were equal across<br />

conditions. Twelve subjects were scanned at 4T (2-shot T2* sensitive<br />

EPI, TR = 1000 ms, TE = 30 ms, flip angle = 20?, 64 x 64 voxel matrix,<br />

192 x 192 mm FOV, in-plane resolution = 3.5 mm, through-plane<br />

resolution = 3.5 mm). Activity in primary olfactory cortex was<br />

significantly higher when sniffing clean air in "task detection" than<br />

when sniffing the same content in "task inhalation" (p < .04). In other<br />

words, activity in human primary olfactory cortex was strongly<br />

dependant on task context. Supported by NSF.<br />

81<br />

311 Slide [ ] Cortical Signal Processing<br />

INFORMATION CODING IN THE OLFACTORY SYSTEM<br />

Buck L. 1, Zou Z. 1 1Basic <strong>Sciences</strong>, Fred Hutchinson Cancer Research<br />

Center, Seattle, WA<br />

Odorants are detected in mice by 1000 different odorant receptors<br />

(ORs) that are expressed by olfactory sensory neurons in the nose. The<br />

ORs are used combinatorially to detect different odorants and encode<br />

their unique identities. To explore the mechanisms underlying odor<br />

perception, we asked how inputs derived from different mouse ORs are<br />

organized as signals travel from the nose to the olfactory bulb and then<br />

the olfactory cortex. In the nose, each sensory neuron expresses a single<br />

OR gene. Neurons with the same OR are dispersed in the nose, but their<br />

axons converge in a few glomeruli at two fixed locations in the bulb.<br />

The result is a stereotyped sensory map in which inputs from different<br />

ORs are segregated in different glomeruli and relay neurons. In the<br />

olfactory cortex, inputs from one OR are targeted to clusters of neurons<br />

at specific sites, creating a stereotyped map unrelated to that in the bulb.<br />

In contrast to the segregation of different OR inputs seen in both the<br />

nose and bulb, it appears that different OR inputs overlap extensively in<br />

the cortex and single neurons may receive combinatorial inputs from<br />

many different ORs. Using c-Fos as an indicator of neuronal activator,<br />

we found that different odorants elicit different, but partially<br />

overlapping, activation patterns in the cortex. The representation of<br />

each odorant is composed of a small subset of sparsely distributed<br />

neurons. Quantitative analysis of the odor representations suggests that<br />

cortical neurons may function as coincidence detectors that are<br />

activated only by correlated inputs from different ORs.<br />

312 Symposium [ ] Chemical Communication in Mammals:<br />

From Pheromones to Individual Recognition<br />

THE MAMMARY PHEROMONE OF THE RABBIT: IDENTITY,<br />

SOURCE, AND SOME FUNCTIONS<br />

Schaal B. 1, Coureaud G. 1, Moncomble A. 1, Langlois D. 2 1Centre des<br />

<strong>Sciences</strong> du Goût, CNRS, Dijon, France; 2UMRA, Inra, Dijon, France<br />

Mammalian females have evolved odor cues to guide their newborns<br />

to their mammae, whereas newborns have coevolved reliable means to<br />

respond to them efficiently. The domestic rabbit is a particularly<br />

suitable model to understand these cues because of the extremely<br />

sparing nature of maternal care en<strong>for</strong>cing nose-lead pups to<br />

instantaneously seize a nipple. Using a GC-olfaction assay on the<br />

volatiles of rabbit milk, an active compound was identified as 2methylbut-2-enal.<br />

It releases stereotypical searching and oral grasping<br />

in pups as effectively as milk itself. Its activity is concentrationdependent<br />

and qualitatively selective as 40 other odorants from milk,<br />

from rabbit secretions or chosen arbitrarily revealed ineffective. The<br />

compound qualifies as a pheromone in the sense defined by Beauchamp<br />

et al (1976) and Johnston (2000) <strong>for</strong> mammals: It triggers invariant<br />

responses of clear functional significance, generalizes across pups of<br />

various genetic and dietary backgrounds, and acts instantly as an<br />

unconditional stimulus in pups devoid of prior exposure with it. As it<br />

seems to be emitted de novo in the mammary tract, we named it<br />

'mammary pheromone' (MP). It is one of the key compounds of ejected<br />

milk as its concentration in milk correlates with behavioural activity. In<br />

more functional terms, the MP has a powerful activating impact on<br />

pups resting in the nest, and directs searching as efficiently as odor cues<br />

emitted on the abdominal surface by lactating does. This predisposed<br />

odour-behavior coupling will be discussed in the context of motheryoung<br />

communication and of adaptive development of offspring in<br />

mammals.


313 Symposium [ ] Chemical Communication in Mammals:<br />

From Pheromones to Individual Recognition<br />

MAKING “SCENTS” OF OWNERSHIP<br />

Hurst J. 1 1Veterinary Clinical Science, University of Liverpool, Wirral,<br />

United Kingdom<br />

In order that a receiver can match in<strong>for</strong>mation in a scent mark to an<br />

individual owner, the signal must provide stable and persistent<br />

in<strong>for</strong>mation about the owner´s individual identity. In mice (Mus<br />

domesticus), two polymorphic gene complexes contribute to the scents<br />

that differ between individuals: the major histocompatibility complex<br />

(MHC) and the major urinary proteins (MUPs). These polymorphic and<br />

polygenic proteins are “hard-wired” in the genome and could provide<br />

stable ownership signals, while volatiles determined by developmental<br />

and environmental as well as genetic factors provide in<strong>for</strong>mation on the<br />

current status of the owner. MHC-associated odours are important in<br />

promoting MHC-disassortative mate selection and kin recognition, but<br />

their contribution to individual identity signatures is unclear. Recent<br />

experiments reveal that individual MUP patterns are the basis of<br />

ownership signals in male competitive scent marks. Our current<br />

hypothesis, of an associative mechanism, attempts to bring together the<br />

different contributions made by MHC and MUPs. Volatile components<br />

perceived at a distance induce investigation through contact. The<br />

receiver then learns an association between the variable volatile<br />

signature detected through the main olfactory system and the involatile<br />

MUP signature detected via the vomeronasal system. Plasticity in this<br />

association means that changes in volatile signatures can be reassociated<br />

with stable involatile signatures during contact investigation.<br />

This permits rapid and distant identification when volatile signatures<br />

are familiar, together with a reliable signature of identity even when<br />

status or environment changes. Funded by BBSRC, UK<br />

314 Symposium [ ] Chemical Communication in Mammals:<br />

From Pheromones to Individual Recognition<br />

WHY MUSTH (AND OTHER) ELEPHANTS USE<br />

PHEROMONES<br />

Rasmussen L.E. 1, Greenwood D.R. 2 1OGI School of Science &<br />

Engineering, OHSU, Beaverton, OR; 2Gene Technologies,<br />

HortResearch, Auckland, New Zealand<br />

In the wild Elephas maximus and Loxodonta africana societies are<br />

generally stable and often resilient, two attributes rein<strong>for</strong>ced by<br />

longevity, multigenerational matriarchal families, and multimodal<br />

signaling between individuals. In Asian elephants successful life<br />

strategies and complex behaviors are interlocked with an extensive<br />

chemical communication system. Initially, well-defined behaviors of<br />

wild and captive elephants <strong>for</strong>med the foundation <strong>for</strong> our investigation<br />

of olfactory communication. Two chemically identified pheromones,<br />

(Z)-7-dodecenyl acetate, released by preovulatory females, and<br />

frontalin, released by older adult males in musth, elicit distinctive,<br />

quantitatively measurable and statistically valid sex- and sexual-statespecific<br />

behaviors. The vast size and clearly spatially and functionally<br />

differentiated anatomical interrelationships of the main olfactory and<br />

vomeronasal organ systems not only allow the investigation of systems<br />

interplay, but effectively slow down individual biochemical events as<br />

temporally discrete steps, allowing such biochemical olfactory events to<br />

be correlated in real time with behavioral actions. The partially<br />

elucidated pathways <strong>for</strong> pheromone transport from trunk tip to sensory<br />

epithelia involving specific binding proteins will be demonstrated using<br />

concurrent biochemical and behavioral experiments, and will include<br />

data on protein-pheromone interactions and kinetic analyses,<br />

demonstrating that our Asian elephant model of vertebrate olfaction is<br />

an attractive experimental system with which to study distinctive steps,<br />

especially periceptive events, in discrete time.<br />

82<br />

315 Symposium [ ] Chemical Communication in Mammals:<br />

From Pheromones to Individual Recognition<br />

INDIVIDUAL RECOGNITION: SIGNALS, BEHAVIOR AND<br />

NEURAL MECHANISMS<br />

Johnston R.E. 1 1Psychology, Cornell University, Ithaca, NY<br />

Recognition of individuals is essential <strong>for</strong> many types of social<br />

behavior and reproductive fitness in vertebrates. The cues <strong>for</strong> such<br />

recognition are complex and multi-dimensional; odor cues <strong>for</strong><br />

individual recognition consist of complex mixtures whose proportions<br />

vary across individuals (mosaic signals). Since other individuals can<br />

have great emotional significance, individually distinctive signals can<br />

have powerful influences on behavior (e.g., fear specific to a familiar<br />

opponent). Hamsters (Mesocricetus auratus) use at least 4 sources of<br />

body odor <strong>for</strong> recognition of individuals and they have integrated,<br />

multi-component memories of such individuals. They also have<br />

extraordinary abilities to analyze scent over-marks and to determine<br />

which individual´s scent is on top of the other; they use this in<strong>for</strong>mation<br />

to evaluate the vigor and quality of potential mates. At the neural level<br />

of analysis, both the main olfactory and vomeronasal systems<br />

contribute to discrimination of individuals by odors. Brain imaging<br />

studies using immediate early genes as markers <strong>for</strong> cell activity indicate<br />

circuits that are probably involved in recognition, memory, and<br />

appropriate emotional responses to familiar individuals (specifically,<br />

winners of fights with the subjects). These areas include parts of the<br />

hippocampus (dorsal CA1 area and subiculum), other para-hippocampal<br />

areas, and the basolateral amygdala. This model system provides insight<br />

into the functions of higher-order olfactory processing and areas of the<br />

brain involved in social behavior.<br />

Supported by NIH grant 5R01 MH58001-01A1.<br />

316 Poster [ ] Taste: Animal Behavior<br />

MICROSTRUCTURAL ANALYSIS OF LICKING IN THE<br />

FORMATION AND EXTINCTION OF A CONDITIONED<br />

TASTE AVERSION<br />

Baird J.P. 1, St. John S.J. 2, Nguyen E.A. 1 1Psychology, Amherst College,<br />

Amherst, MA; 2Reed College, Portland, OR<br />

LiCl (ip) paired with intraoral sucrose causes sucrose taste reactivity<br />

to shift to a profile that resembles aversive intraoral QHCl. We used<br />

lick microstructure analysis to assess changes in oromotor activity<br />

under conditions that may better approximate natural toxin exposure.<br />

Water-deprived rats were trained to access dh2o 15 min/day from a<br />

bottle connected to a lickometer. On days 1,3,&5 0.12M LiCl (8 rats) or<br />

0.12M NaCl (8 rats) solution replaced dh2o. On days 7,9,&11 only<br />

0.12M NaCl replaced dh2o. A 2nd study used 0.24M NaCl instead of<br />

0.12M NaCl. On the 1st LiCl day rats avidly drank LiCl <strong>for</strong> 3 min: 1st<br />

min lick rates and mean burst durations were comparable to NaCl<br />

controls. As LiCl licking progressed, ingestion rate slowed significantly<br />

by the 4th-6th minute vs. controls. On subsequent LiCl tests, intake,<br />

initial rate, burst duration, ingestion rate & lick volume were<br />

significantly reduced. However, burst count was significantly increased.<br />

In rats ingesting NaCl after LiCl, intake, initial lick rate (except .24M<br />

NaCl), ingestion rate, burst duration & lick volume remained<br />

significantly reduced relative to controls, and burst count remained<br />

increased by at least 72%. Over remaining NaCl tests effects reversed<br />

and almost all differences were extinguished by the last NaCl day.<br />

These effects of LiCl parallel microstructural responses to 0.2mM<br />

QHCl, and shifts in taste reactivity after CTA <strong>for</strong>mation. The generality<br />

of these results across two paradigms supports the notion that CTA´s<br />

can affect hedonic evaluation of tastants. The construct validity <strong>for</strong> both<br />

techniques as useful tools in the measurement of hedonic gustatory<br />

processes is also supported. Supported by a Mellon-8 grant to JPB.


317 Poster [ ] Taste: Animal Behavior<br />

D-CYCLOSERINE POTENTIATES SHORT-DELAY, BUT NOT<br />

LONG-DELAY, CONDITIONED TASTE AVERSION.<br />

Davenport R.A. 1, Houpt T.A. 1 1Biological <strong>Sciences</strong>, Florida State<br />

University, Tallahassee, FL<br />

D-Cycloserine (DCS) is a partial glycine agonist at the NMDA<br />

glutamate receptor site, and has been shown to facilitate both<br />

declarative and non-declarative memory tasks. In order to determine if<br />

DCS enhances conditioned taste aversion (CTA) learning, we<br />

administered DCS be<strong>for</strong>e the pairing of saccharin intake and LiCl<br />

injection and measured expression with 2-bottle tests.<br />

Water-deprived rats were injected with DCS (15mg/kg i.p.) 15 min<br />

prior to 10-min access to 0.125% saccharin. Rats were injected with<br />

LiCl (19mg/kg i.p.) 25 or 60 min after DCS. Controls were injected<br />

with saline in place of DCS, or saline in place of LiCl. One day after<br />

conditioning, rats were given 24-h, 2-bottle preference tests <strong>for</strong> 14 days.<br />

Daily saccharin preference was calculated as saccharin intake over total<br />

intake.<br />

Control rats showed no CTA when saccharin was paired with saline<br />

(pref: 0.9 ± 0.04), and only a moderate CTA after LiCl injection in the<br />

absence of DCS (pref: 0.4 ± 0.1, p < 0.01 vs. saline). In the presence of<br />

DCS, rats showed a stronger aversion after LiCl at 25 min (pref: 0.06 ±<br />

0.02, p < 0.05 vs. LiCl alone), but the same aversion after LiCl at 60<br />

min (pref: 0.32 ± 0.1).<br />

We conclude that DCS enhances a CTA but is dependent on the<br />

timing of both the gustatory stimulus and the toxin, such that DCS<br />

potentiated short-delay but not long-delay pairing of saccharin and<br />

LiCl. This difference may be due to the short-half life of DCS, or it may<br />

reflect different roles <strong>for</strong> NMDA neurotransmission in gustatory and<br />

toxin processing. Current studies are probing different time points with<br />

DCS to distinguish these possibilities. Supported by an FSU<br />

Neuroscience Fellowship and NIDCD 03198.<br />

83<br />

318 Poster [ ] Taste: Animal Behavior<br />

DIFFERENCES IN GUSTATORY BEHAVIOR BETWEEN<br />

C57BL/6J AND DBA/2J INBRED MICE<br />

Raghow S. 1, Boughter J.D. 1, Nelson T.M. 2, St. John S.J. 3, Munger S.D. 2<br />

1Anatomy & Neurobiology, University of Tennessee, Memphis, TN;<br />

2Anatomy and Neurobiology, University of Maryland at Baltimore,<br />

Baltimore, MD; 3Reed College, Portland, OR<br />

The commonly used C57BL/6J (B6) and DBA/2J (D2) strains of<br />

inbred mice are among the strains included in the public and private<br />

genome sequencing projects and, given the availability of BXD<br />

recombinant inbred (RI) strains, provide an extremely desirable model<br />

<strong>for</strong> genetic studies. B6 and D2 mice have been shown to differ in terms<br />

of intake behavior to taste stimuli of different qualities, illustrating the<br />

potential of the BXD RI set <strong>for</strong> mapping QTLs influencing taste<br />

sensitivity. However, 24- or 48-hr intake tests are generally confounded<br />

by non-gustatory or post-ingestive factors such as olfactory cues,<br />

caloric load, and toxicity. We examined gustatory behavior, especially<br />

to an array of stimuli characterized as bitter or aversive, in B6 and D2<br />

mice using both intake and brief-access tests. Robust strain differences<br />

in sensitivity <strong>for</strong> bitter stimuli QHCl, PROP and RUA were found using<br />

both assays. However, the strains did not differ <strong>for</strong> cycloheximide,<br />

strychnine, or MgCl2 in the brief-access test. We also failed to detect a<br />

strain difference <strong>for</strong> salts using either assay. We also developed a 4-day,<br />

high-throughput screen <strong>for</strong> bitter taste sensitivity. B6, D2, and BXD F1<br />

and F2 intercross mice were tested with two concentrations each of<br />

denatonium benzoate, PROP, and QHCl. Mice were also assessed <strong>for</strong><br />

additional non-gustatory phenotypes, including water lick rate, water<br />

consumption, latency to first lick and overall per<strong>for</strong>mance in the task.<br />

B6 and D2 mice significantly differed in taste sensitivity <strong>for</strong> PROP and<br />

QHCl. However, sensitivity to PROP and QHCL were not correlated (r<br />

= 0.08) in 126 F2 mice, demonstrating independence of these<br />

phenotypes. Supported by NIDCD: DC005786(SDM),<br />

DC004935(JDB).<br />

319 Poster [ ] Taste: Animal Behavior<br />

GAPING TO QUININE IN GLOSSOPHARYNGEAL NERVE-<br />

TRANSECTED RATS AFTER POSTSURGICAL TASTE<br />

AVERSION CONDITIONING<br />

Bayevsky A. 1, Colbert C.L. 1, Garcea M. 1, Newth A. 1, Spector A.C. 1<br />

1Psychology Department, University of Florida, Gainesville, FL<br />

Glossopharyngeal nerve transection (GLX) in rats has been shown to<br />

significantly reduce gaping, a stereotypical oromotor rejection behavior,<br />

in response to intraoral quinine hydrochloride (Q) infusion. We<br />

examined whether gaping to Q could be increased by a conditioned<br />

taste aversion (CTA) to this stimulus in GLX rats. Rats had intraoral<br />

cannulae implanted and either received sham surgery (SHAM) or GLX.<br />

After habituation to the test chamber, animals were infused with 0.3 M<br />

sucrose <strong>for</strong> 3 minutes (1 mL/min) on 3 consecutive days, followed by<br />

0.6 mM Q infusion, the latter of which preceded an injection (inj) of<br />

LiCl or NaCl (2.0 mEq/Kg). The rats were videorecorded and this 4-day<br />

cycle was repeated 3 more times with no injections given after the last<br />

Q infusion (test day). The first 30 s of the test-day Q infusion were<br />

scored <strong>for</strong> gapes adjusted <strong>for</strong> unscorable periods. The LiCl-inj SHAM<br />

(n=7) and GLX (n=9) groups gaped significantly more to Q compared<br />

with the NaCl-inj GLX (n=7) rats and did not differ from each other.<br />

Gaping in the NaCl-inj SHAM (n=8) rats fell between these two<br />

extremes and only significantly differed from that in the LiCl-inj GLX<br />

group. The lack of an expected significant difference in gapes between<br />

the 2 NaCl-inj groups could be due to repeated stimulus exposure<br />

effects. We are scoring the first Q infusion to evaluate this and to allow<br />

us to conduct within-subjects analyses. At present, these findings<br />

suggest that a CTA can increase gaping to Q in GLX rats as has been<br />

previously shown <strong>for</strong> sucrose. Supported by NIH R01-DC01628.


320 Poster [ ] Taste: Animal Behavior<br />

TASTE PREFERENCE AND TASTE BUDS MAINTENANCE<br />

AFTER UNILATERAL LINGUAL DENERVATION IN RATS.<br />

Lee J. 1, Kim Y. 1, Moon Y. 2, Jahng J. 3 1Oral & Maxillofacial Surgery,<br />

Seoul National University College of Dentistry, Seoul, South Korea;<br />

2Biology, Catholic University College of Medicine, Seoul, South Korea;<br />

3Yonsei University College of Medicine, Seoul, South Korea<br />

Intact gustatory nerve supply appears to be required <strong>for</strong> the<br />

maintenance of taste buds, and unilateral transection of gustatory nerve<br />

results in reduced number of taste papillae with degeneration and<br />

disappearance of taste buds ipsilateral to the surgical side. We<br />

examined the effects of unilateral lingual nerve transection on taste<br />

function as well as on the maintenance of taste buds. Male Sprague-<br />

Dawley rats weighing 250-300g received unilateral transection of<br />

lingual nerve, subjected to the preference test <strong>for</strong> various taste solutions<br />

(0.1M NaCl, 0.1M sucrose, 0.01M QHCl, or 0.01M HCl) with two<br />

bottle test paradigm at 2, 4, 6, or 8 weeks after the operation. Water and<br />

each test solution bottle were supplied <strong>for</strong> 48 h at each test with one<br />

change of the bottle position at 24 h. The preference score <strong>for</strong> salty,<br />

sweet or sour taste, but not <strong>for</strong> bitter taste, tended to be higher in the<br />

operated rats without statistical significance, compared to the sham rats.<br />

Overall pattern of the preference scores was not changed by time after<br />

operation. These results suggest that unilateral damage of lingual nerve<br />

may not significantly affect taste function of the subject. In order to<br />

examine the morphologic changes of tongue, number of fungi <strong>for</strong>m<br />

papillae on the denervated side was counted in comparison with the<br />

intact side. Time course of apoptotic death of taste cells in the foliate<br />

and vallate papillae was examined as well. Supported by the Korea<br />

Health R & D Project (JHL), KISTEP(JWJ).<br />

321 Poster [ ] Taste: Peripheral Connectivity<br />

ULTRASTRUCTURE OF MORPHOLOGICALLY IDENTIFIED<br />

CHORDA TYMPANI AXONS IN THE NUCLEUS OF THE<br />

SOLITARY TRACT IN DEVELOPMENTALLY SODIUM-<br />

RESTRICTED AND CONTROL RATS<br />

May O.L. 1, Erisir A. 1, Hill D.L. 1 1Psychology, University of Virginia,<br />

Charlottesville, VA<br />

Dietary sodium restriction during a critical period in gestation results<br />

in specific, yet profound morphological changes in the dorsal zone of<br />

the chorda tympani nerve terminal field in the adult rat nucleus of the<br />

solitary tract. Specifically, there is an approximate two-fold increase in<br />

volume of the dorsal zone of the terminal field in sodium-restricted rats<br />

compared to controls. In order to better characterize this plasticity, we<br />

examined the ultrastructure of terminals identified by way of bulklabeling<br />

the chorda tympani nerve with biotinylated dextran and<br />

visualizing with DAB. Quantitative measurements in the dorsal zone of<br />

the chorda tympani field in both sodium-restricted and control rats<br />

included volumetric density of labeled synapses and synapsing<br />

frequency of axons. Preliminary evidence indicates that there is over a<br />

four-fold increase in the density of chorda tympani terminals in the<br />

dorsal region in restricted rats compared to controls. However, there<br />

was not a significant increase in synapse density along any given axon,<br />

suggesting that elaboration of chorda tympani input to this region<br />

occurs by addition of new axonal arbors. Furthermore, there is a<br />

conspicuous qualitative difference in the morphology of synaptic<br />

terminals in sodium-restricted rats in that their profiles are interrupted<br />

by finger-like protrusions of dendritic spines. These results indicate a<br />

process of synaptic remodeling as well as an elaboration of connectivity<br />

occurs as a result of developmental sodium restriction. Supported by<br />

NIH grants DC00407 and F31 DC06332<br />

84<br />

322 Poster [ ] Taste: Peripheral Connectivity<br />

GUSTATORY NERVE TERMINAL FIELDS IN RATS<br />

RECOVERED FROM EARLY DEVELOPMENTAL SODIUM<br />

RESTRICTION.<br />

Mangold J.E. 1, Hill D.L. 1 1Psychology, University of Virginia,<br />

Charlottesville, VA<br />

A critical period <strong>for</strong> chorda tympani (CT) terminal field development<br />

is from embryonic day 3 (E3) to E12. When pregnant rats are fed a lowsodium<br />

diet during this period, the CT terminal field in their offspring<br />

(E3-E12 restricted) is enlarged compared to controls. Similarly, rats<br />

raised on a sodium-restricted diet from E3 to postnatal day 28 and then<br />

fed a sodium-replete diet <strong>for</strong> at least a month (P28 recovered) also have<br />

CT terminal fields that are large. To determine the interactions among<br />

the CT, GSP and IX terminal fields in controls, E3-E12 restricted, and<br />

P28 recovered rats, we used a triple fluorescent labeling procedure. In<br />

control rats, the IXth field is the most dorsal of the three fields and<br />

extends ventrally into the dorsal-most zone of the GSP. The CT field<br />

overlaps much of the GSP field, but has no overlap with the IXth field.<br />

In E3-E12 sodium restricted rats, the IXth field is approximately 3X<br />

greater than that in controls and extends medio-laterally and ventrally,<br />

well into the CT field. Both the GSP and CT fields are significantly<br />

larger than controls. P28 recovered rats had similar alterations of all<br />

three fields; however, they were not as dramatic as in E3-E12 restricted<br />

rats. There<strong>for</strong>e, a very early period of dietary sodium restriction (E3-<br />

E12) that precedes the development of peripheral gustatory structures<br />

produces the most widespread effects on the development of gustatory<br />

afferents in the NTS. These results also suggest a greater functional<br />

convergence in the gustatory brainstem and behavioral consequences<br />

<strong>for</strong> sodium restriction during development. Supported by grant 00407.<br />

323 Poster [ ] Taste: Peripheral Connectivity<br />

ISOFORMS OF THE SYNAPTIC VESICLE PROTEIN SV2<br />

HAVE DIFFERENT LOCATIONS IN THE RAT<br />

CIRCUMVALLATE GUSTATORY TISSUE<br />

Nelson G.M. 1 1Neurobiology, University of Alabama at Birmingham,<br />

Birmingham, AL<br />

SV2, a 12-transmembrane protein which may function as a calcium<br />

regulator in neurotransmission and a component of the intravesicular<br />

matrix, has three identified iso<strong>for</strong>ms, A, B, and C. Presynaptic size<br />

vesicles are present in the gustatory system in primary afferent nerve<br />

fibers receiving synapses from taste cells, peptidergic intragemmal<br />

nerve fibers, and taste cells. Two color fluorescent microscopy of fixed<br />

rat tongue tissue sections is labeled with various SV2 antibodies. A<br />

general SV2 antibody labels most nerve fibers, including intragemmal,<br />

some perigemmal, peri-apical pore, basal plexus, and the sub-epithelial<br />

network of nerve fibers. SV2B has a very restricted pattern suggestive<br />

of location in taste cells, especially in taste buds in the lower<br />

circumvallate crypt. SV2C labels a smaller group of intragemmal,<br />

perigemmal, and basal plexus fibers than the general SV2 antibody. The<br />

SV2 general positive fibers are included in the population of fibers<br />

which label with PGP 9.5. Most SV2B and C positive structures<br />

colocalize with PGP-positive structures. In contrast, a few intragemmal<br />

fibers labeled with either the general SV2 or SV2B antibody appear to<br />

colocalize or oppose gustducin-positive taste cells. These results<br />

indicate that the antibodies to various SV2 iso<strong>for</strong>ms may be a tool to<br />

differentiate various pools of small clear vesicles within the gustatory<br />

system. SV2-positive nerve fibers are a subset of the PGP-positive<br />

fibers and may identify synaptic sites between some nerve fibers and<br />

taste cells. Supported by NIDCD 00166.


324 Poster [ ] Taste: Peripheral Connectivity<br />

SYNAPTOPHYSIN-LIKE IMMUNOREACTIVITY IN<br />

CIRCUMVALLATE PAPILLAE OF THE RAT AND MOUSE<br />

Schmidt K.C. 1, Yang R. 2, Kinnamon J.C. 3 1Biological <strong>Sciences</strong>,<br />

University of Denver, Denver, CO; 2University of Denver, Denver, CO;<br />

3Biology, University of Denver, Denver, CO<br />

Synaptophysin is an abundant protein in the CNS that is found in the<br />

membranes of synaptic vesicles. Synaptophysin is thought to play a role<br />

in the regulation of SNARE <strong>for</strong>mation during synaptic transmission<br />

through its interactions with other proteins of the synaptic vesicle cycle,<br />

namely binding with VAMP/synaptobrevin. We postulate that<br />

synaptophysin is associated with the synaptic vesicles at taste cell<br />

synapses. Our preliminary results indicate that subsets of taste cells of<br />

the circumvallate papillae of both the rat and mouse display<br />

synaptophysin-like immunoreactivity (LIR). The immunoreactive cells<br />

are slender and span the entire length of the taste bud from the basal<br />

lamina to the taste pore. In the rat, nerve processes exiting the taste<br />

buds are also immunoreactive. Synaptophysin-LIR does not appear to<br />

colocalize with IP3R3-LIR, but does appear to colocalize with a subset<br />

of PGP 9.5 (protein gene product 9.5) immunoreactive cells. The<br />

elucidation of proteins such as synaptophysin involved in the synaptic<br />

vesicle cycle in taste cells may prove useful in determining how taste<br />

cells transmit in<strong>for</strong>mation. Supported by NIH grants DC00285 and<br />

DC00244.<br />

325 Poster [ ] Taste: Peripheral Connectivity<br />

IMMUNOCYTOCHEMICAL ANALYSIS OF SYNTAXIN IN<br />

RAT CIRCUMVALLATE TASTE BUDS<br />

Yang R. 1, Thomas S. 1, Kinnamon J.C. 1 1Biological <strong>Sciences</strong>, University<br />

of Denver, Denver, CO<br />

The SNARE proteins syntaxin, SNAP-25, and synaptobrevin all play<br />

key roles during Ca2+ dependent exocytosis in the CNS. We<br />

hypothesize that taste cell synapses utilize the same protein machinery<br />

as used by synapses in the CNS. Our previous studies have shown that<br />

taste cells with synapses display SNAP-25- and synaptobrevin-like<br />

immunoreactivity (LIR) (Yang et al., 2000; Yang et al., in press). At<br />

present, we are studying the presynaptic membrane protein, syntaxin, in<br />

rat circumvallate taste buds. Our current results indicate that syntaxin is<br />

present in a subset of taste cells and nerve processes in taste buds.<br />

Approximately 15% of the taste cells in a taste bud display cytoplasmic<br />

syntaxin-LIR and a smaller subset of taste cells also show punctate<br />

staining in the cytoplasm. Syntaxin-LIR colocalizes with SNAP-25- and<br />

synaptobrevin-LIR in a small subset of taste cells and most nerve<br />

processes. Approximately 11% of the syntaxin-LIR cells colocalize<br />

with PLCβ2-LIR cells. However, only about 4% of the PLCβ2-LIR<br />

cells also display syntaxin-LIR. DAB immunoelectron microscopy<br />

reveals that syntaxin-LIR is present in both type II and III taste cells.<br />

Type II taste cells display punctate syntaxin-LIR at Golgi bodies, while<br />

type III taste cells show cytoplasmic syntaxin-LIR. All of the synapses<br />

associated with syntaxin-LIR taste cells are from type III cells onto<br />

nerve processes. These results support the notion that taste cell synapses<br />

are similar to synapses of the CNS. Supported by NIH grant DC00285.<br />

85<br />

326 Poster [ ] Taste: Peripheral Connectivity<br />

TASTE BUDS AND SURROUNDING FIBERS ARE<br />

IMMUNOREACTIVE FOR THE IONOTROPIC ATP<br />

RECEPTOR P2X7<br />

Stone L.M. 1, Kinnamon S.C. 1 1Biomedical <strong>Sciences</strong>, Colorado State<br />

University, Fort Collins, CO<br />

Processing of taste in<strong>for</strong>mation probably involves intercellular<br />

communication within the taste bud as well as communication between<br />

individual taste cells and innervating nerve fibers.<br />

Immunocytochemical studies of transmitter proteins and receptors<br />

suggest that several neurotransmitter pathways are used by taste buds.<br />

Recent evidence suggests a role <strong>for</strong> ATP and its receptors in taste bud<br />

function (Bo et al., 1999; Y.V. Kim et al., 2000; Baryshnikov et al.,<br />

2003). Further, physiological studies suggest taste cell responses to<br />

ATP involve metabotropic (P2Y-like) receptors. The current study was<br />

done to determine if ionotropic ATP receptors also are present on taste<br />

cells. Tongues from transgenic mice expressing GFP under the control<br />

of the α-gustducin promoter were examined <strong>for</strong> the presence of P2X7<br />

immunoreactivity. We found that P2X7 immunoreactivity includes taste<br />

cells as well as nerve fibers in and around taste buds. P2X7 occurs both<br />

in cells that express α-gustducin and in taste cells that lack the G<br />

protein. In conclusion, taste cells may signal to one another by both<br />

metabotropic and ionotropic ATP receptors. Supported by DC00766<br />

327 Poster [ ] Taste: Peripheral Connectivity<br />

RECOVERY OF GURMARIN-SENSITIVE NEURAL<br />

RESPONSES AND EXPRESSION OF T1R3 AND GUSTDUCIN<br />

IN FUNGIFORM PAPILLAE AFTER CRUSH OF THE MOUSE<br />

CHORDA TYMPANI<br />

Yasumatsu K. 1, Shigemura N. 1, Shigeoka Y. 1, Ninomiya Y. 1 1Oral<br />

Neuroscience, Kyushu University, Fukuoka, Japan<br />

Gurmarin is known to be a peptide that selectively inhibits responses<br />

to sweet compounds in the chorda tympani nerve (CT) in C57BL mice.<br />

The peptide, however, suppresses only about half of the response to<br />

sucrose, indicating existence of both gurmarin-sensitive (GS) and -<br />

insensitive (GI) response components in the CT. In the present study,<br />

recovery of taste responses and reappearance of taste receptor cells<br />

were studied by examining GS of the CT responses and expression of<br />

T1R3 and gustducin in fungi<strong>for</strong>m papillae. It was reported that T1R3<br />

contributes responses to sucrose, saccharin and other artificial<br />

sweeteners in the CT. At about 2 weeks after the nerve crush, although<br />

no significant responses to taste stimuli were observed in the CT, in situ<br />

hybridization analysis demonstrated that T1R3 and gustducin expressed<br />

in subsets of taste bud cells. At about 3 weeks after the CT crush,<br />

responses to sweet compounds reappeared and the threshold of sucrose<br />

was higher than that shown by intact CT. After more than a month, the<br />

CT showed GS comparable with those shown by intact animals. These<br />

data suggest that the molecules related to taste transduction might<br />

express in taste receptor cells prior to their contact with regenerated<br />

taste axons.


328 Poster [ ] Taste: Peripheral Connectivity<br />

THE NEURAL ISOFORM OF TRYPTOPHAN HYDROXYLASE<br />

IS LOCALIZED TO TASTE BUD CELLS<br />

Cao J. 1, Huang L. 1, Brand J. 1 1Monell Chemical Senses Center,<br />

Philadelphia, PA<br />

Taste receptor cells are epithelial in origin and <strong>for</strong>m synaptic contact<br />

with innervating sensory nerves. The identification of the<br />

neurotransmitter(s) in these cells has, however, not been absolutely<br />

determined. While a number of studies implicate serotonin as a<br />

candidate neurotransmitter, evidence of its synthetic pathway in taste<br />

receptor cells is incomplete. In the current study, RT-PCR identified<br />

the neural iso<strong>for</strong>m of tryptophan hyroxylase (TPH2) in rat foliate and<br />

vallate isolated taste buds and single taste bud cells (TBC). Sequencing<br />

of 666 bp near the 3´-end of the taste TPH2 showed this product to have<br />

99% identity with rat brain TPH2. The peripheral iso<strong>for</strong>m, TPH1, could<br />

not be detected by RT-PCR in taste tissue, but was detected in brain<br />

cDNA, used as a positive control. Of 12 TBC wherein the synaptic<br />

marker, syntaxin, was detected by RT-PCR, TPH2, but not TPH1, was<br />

detected in 10 of these cells, and T1R3 detected in 6.<br />

Immunocytochemistry localized TPH2 to a subset of taste bud cells<br />

located within the main body of the bud. No staining was seen in the<br />

basal region nor in the perigemmal cells. The results support a role <strong>for</strong><br />

serotonin as a neurotransmitter in taste receptor cells.<br />

Supported in part by a grant to JGB from the Department of Veterans<br />

Affairs, and by NIH DC05154 to L.H.<br />

329 Poster [ ] Taste: Peripheral Connectivity<br />

TASTE BUDS RELEASE 5HT WHEN DEPOLARIZED<br />

Huang Y.J. 1, Lu K.S. 1, Roper S.D. 2 1Physiology & Biophysics,<br />

University of Miami, Miami, FL; 2Physiology & Biophysics and<br />

Neuroscience Program, University of Miami, Miami, FL<br />

Serotonin (5HT) is implicated as a neurotransmitter or<br />

neuromodulator in taste buds but to date there has been no direct<br />

evidence <strong>for</strong> its release from taste cells. Using a CHO/biosensor cell,<br />

we tested whether isolated mouse vallate taste buds release 5HT in<br />

response to stimulation. Pilot studies confirmed that CHO cells, stably<br />

transfected with 5HT2c receptors (Berg et al, Molec Pharmacol, 1994)<br />

and loaded with the Ca2+ -sensitive dye, Fura2, to assay responses, were<br />

exquisitely sensitive to 5HT (threshold ~3 nM). These "CHO/biosensor<br />

cells" did not respond to bath-applied KCl (50-100 mM) nor to<br />

cycloheximide (100 µM), a bitter taste stimulus. We confirmed that<br />

5HT-responses in CHO/biosensor cells were reversibly inhibited by<br />

mianserin (1-10 nM), a 5HT2c antagonist. We isolated taste buds from<br />

mouse vallate papillae and loaded them with Fura2 to verify that taste<br />

cells responded to KCl depolarization (50-100 mM). Finally, to test<br />

whether stimulated taste buds released 5HT, we isolated individual taste<br />

buds and plated them on a glass coverslip in a shallow recording<br />

chamber. Taste buds were approached with a Fura2-loaded<br />

CHO/biosensor cell held onto a fire-polished micropipette by gentle<br />

suction. Bath-applied KCl elicited responses in the CHO/biosensor cell<br />

when it was positioned against a taste bud. Applying buffer did not<br />

elicit responses. Furthermore, moving the CHO/biosensor away from<br />

the taste bud abolished responses. These data suggest that when they<br />

are depolarized, isolated taste buds release 5HT. Supported by DC 6077<br />

and DC00374 (SDR).<br />

86<br />

330 Poster [ ] Taste: Peripheral Connectivity<br />

ANALYSIS OF A HUMAN FUNGIFORM PAPILLAE CDNA<br />

LIBRARY AND IDENTIFICATION OF TASTE-RELATED<br />

GENES<br />

Rossier O. 1, Cao J. 2, Huque T. 2, Spielman A.I. 3, Feldman R.S. 4,<br />

Medrano J.F. 5, Brand J.G. 2, Le Coutre J. 1 1Nestlé Research Center,<br />

Lausanne, Switzerland; 2Monell Chemical Senses Center, Philadelphia,<br />

PA; 3College of Dentistry, New York University, New York, NY; 4Dental<br />

Medicine, V.A. Medical Center, Philadelphia, PA; 5Department of<br />

Animal Science, University of Cali<strong>for</strong>nia, Davis, CA<br />

Various genes related to early events in human gustation have<br />

recently been discovered, yet a thorough understanding of taste<br />

transduction is hampered by gaps in our knowledge of the signaling<br />

chain. As a first step toward gaining additional insight, the expression<br />

specificity of genes in human taste tissue needs to be determined. To<br />

this end, a fungi<strong>for</strong>m papillae cDNA library has been generated and<br />

analyzed. For validation of the library, taste-related gene probes were<br />

used to detect known molecules. Subsequently, DNA sequence analysis<br />

was per<strong>for</strong>med to identify further candidates. Of 987 clones sequenced,<br />

clustering results in 288 contigs. Comparison of these contigs with<br />

genomic databases reveals that 207 contigs (71.9%) match known<br />

genes, 16 (5.6%) match hypothetical genes, 8 (2.8%) match repetitive<br />

sequences and 57 (19.8%) have no or low similarity to annotated genes.<br />

The results indicate that despite a high level of redundancy, this human<br />

fungi<strong>for</strong>m cDNA library contains specific taste markers and is valuable<br />

<strong>for</strong> investigation of both known and novel taste-related genes.<br />

331 Poster [ ] Human Olfaction: Pathology<br />

OLFACTORY FUNCTIONS AND VOLUMES OF<br />

ORBITOFRONTAL AND LIMBIC REGIONS IN<br />

SCHIZOPHRENIA<br />

Rupp C.I. 1, Fleischhacker W.W. 1, Kemmler G. 1, Kremser C. 2, Bilder<br />

R.M. 3, Mechtcheriakov S. 1, Szeszko P.R. 4, Walch T. 1, Scholtz A.W. 5,<br />

Klimbacher M. 1, Maier C. 1, Albrecht G. 1, Lechner T. 1, Felber S. 2,<br />

Hinterhuber H. 1 1University Clinics of Psychiatry Innsbruck, Innsbruck,<br />

Austria; 2University Clinics of Magnetic Resonance Imaging Innsbruck,<br />

Innsbruck, Austria; 3Geffen School of Medicine at UCLA, Los Angeles,<br />

CA; 4Zucker Hillside Hospital, Glen Oaks, NY; 5University Clinics of<br />

Otorhinolaryngology Innsbruck, Innsbruck, Austria<br />

Olfactory deficits in schizophrenia have been widely reported. These<br />

deficits have been hypothesized to be associated with abnormalities in<br />

prefrontal (orbitofrontal) and/or mesiotemporal brain regions. No study<br />

investigated this structure-function relationship using MR-imaging. We<br />

examined the relationship between olfactory functions and volumes of<br />

limbic (hippocampus/amygdala) and orbitofrontal brain regions in<br />

young men with schizophrenia and healthy subjects, matched <strong>for</strong> sex<br />

and age. Unirhinal assessment of olfactory measures included main<br />

functions (threshold, discrimination, identification) and odor<br />

judgements (intensity, familiarity, edibility, pleasantness). Patients<br />

per<strong>for</strong>med bilaterally more poorly than controls in the threshold,<br />

discrimination and identification tasks, as well as on several odor<br />

judgement measures. Compared with controls patients showed bilateral<br />

smaller hippocampus and amygdala volumes. In patients, smaller<br />

volumes of the hippocampus were significantly correlated with poorer<br />

discrimination per<strong>for</strong>mance. Our results corroborate and extend<br />

previous findings of olfactory deficits as well as limbic structure<br />

volume reduction in schizophrenia, and suggest that olfactory deficits,<br />

namely impairments in discrimination, are associated with<br />

morphometric abnormalities in the hippocampus.<br />

Supported by OENB (no. 6576)


332 Slide [ ] Human Olfaction: Pathology<br />

DIMINISHED POSTERIOR NASAL VOLUMES IN MALE<br />

PATIENTS WITH SCHIZOPHRENIA<br />

Roalf D.R. 1, Turetsky B.I. 2, Balderston C.C. 1, Gur R.E. 1, Moberg P.J. 2<br />

1Schizophrenia Research Center, Department of Psychiatry, University<br />

of Pennsylvania, Philadelphia, PA; 2Schizophrenia Research Center,<br />

Department of Psychiatry & Smell & Taste Center, Department of<br />

Otorhinolaryngology:, University of Pennsylvania, Philadelphia, PA<br />

Background: Gross midline abnormalities such as cleft palate and<br />

cavum septum pelucidum, impairments in olfactory ability and to a<br />

lesser degree subtle craniofacial dysmorphogenesis are considered<br />

characteristics of schizophrenia. Due to concordant development of the<br />

oral cavity, olfactory structures, and ventral <strong>for</strong>ebrain, we hypothesized<br />

that structural abnormalities of the nasal cavity might represent diseaseassociated<br />

markers of embryological dysmorphogenesis.<br />

Method: A measurement of nasal volume was acquired by acoustic<br />

rhinometry <strong>for</strong> 56 schizophrenia patients (44 men, 12 women) and 37<br />

healthy controls (24 men, 13 women). Nasal volume was segmented<br />

into three areas: Total (0.0-5.5cm), Anterior (0.0-3.5cm) and Posterior<br />

(3.5-5.5cm).<br />

Results: An overall MANOVA demonstrated a significant diagnosis<br />

by sex by nasal compartment interaction (F[2,172]=2.92, p=.05).<br />

Decomposition of this interaction revealed that male patients had<br />

significantly smaller left (F[2,126]= 6.26, p0.05 on Student's twotailed<br />

t test. Mean UPSIT score <strong>for</strong> controls was 33.5 and ET 32.2<br />

which is not significantly different. For IPD the mean score was 18.1<br />

which is significantly different from controls and ET. OERP showed no<br />

Difference <strong>for</strong> P2 latency between controls and ET patients (mean<br />

latency 628.5ms and 641.3ms respectively). In 21 PD patients the<br />

evoked response was absent. In the remaining 33 the mean latency was<br />

962ms which is significantly prolonged compared to controls and ET.<br />

In all three groups amplitude measurements did not differ significantly.<br />

A normosmic patient with tremor is more likely to have ET than IPD<br />

while someone with tremor and impaired olfaction may have IPD or<br />

related syndrome.<br />

334 Poster [ ] Human Olfaction: Pathology<br />

MEMORY FOR EMOTIONAL AND NEUTRAL ODORS AND<br />

AMYGDALA ELECTROPHYSIOLOGICAL RECORDINGS IN<br />

PATIENTS WITH EPILEPSY<br />

Julie H. 1, Sandra P. 1, Jean G. 1, Marilyn J. 1 1Neuropsychology and<br />

Cognitive Neuroscience, McGill University, Montreal, Quebec, Canada<br />

It is known that emotionally arousing events are remembered better<br />

than neutral ones, but this has not been explored in olfaction. Increasing<br />

evidence indicates that the amygdala plays key roles in emotion and<br />

memory, and electrophysiological studies show that the human<br />

amygdala is implied in recognition memory <strong>for</strong> odors. In the present<br />

study we assessed memory <strong>for</strong> emotional and neutral odors in patients<br />

with temporal-lobe epilepsy undergoing intracranial recordings and<br />

examined the related olfactory evoked potentials (OEPs) in amygdala.<br />

Behavioral and electrophysiological results were obtained in 8 patients<br />

implanted in the amygdala (6 bilaterally, 1 left, 1 right). Patients´<br />

behavioral data were compared to those of 10 healthy control subjects.<br />

Memory <strong>for</strong> pleasant, unpleasant or non emotional odors was tested 24<br />

hours after incidental encoding. Odor recognition was similar in<br />

patients and controls. OEPs were observed in the amygdala, composed<br />

mainly of a large positive peak occurring at a mean latency of 300ms.<br />

An effect of odorant´s emotional arousal was found in the right<br />

amygdala, with shorter latencies <strong>for</strong> emotional than <strong>for</strong> neutral<br />

odorants. This effect was specific to the amygdala, as similar OEPs<br />

were not found in the hippocampus. The results provide<br />

electrophysiological evidence <strong>for</strong> the postulated role of amygdala in<br />

memory <strong>for</strong> emotionally arousing material. The significance of the<br />

laterality of this effect will be explored with respect to the healthiness<br />

of left versus right amygdala in individual patients. Supported by Savoy<br />

Foundation and Canadian Institutes of Health Research.


335 Poster [ ] Human Olfaction: Pathology<br />

OLFACTORY DYSFUNCTION IN DEGENERATIVE ATAXIAS.<br />

Connelly T. 1, Farmer J.M. 2, Lynch D.R. 3, Tourbier I.A. 4, Doty R.L. 4<br />

1Smell and Taste Center, Department of Otorhinolaryngology,<br />

University of Pennsylvania Medical Center, Phila, PA; 2Division of<br />

Medical Genetics, University of Pennsylvania Medical Center,<br />

Philadelphia, PA; 3Neurology, University of Pennsylvania Medical<br />

Center, Philadelphia, PA; 4Smell and Taste Center, Department of<br />

Otorhinolaryngology, University of Pennsylvania Medical Center,<br />

Philadelphia, PA<br />

Several lines of evidence suggest that the cerebellum may play a role<br />

in higher-order olfactory processing. In this study, we administered the<br />

University of Pennsylvania Smell Identification Test (UPSIT), a<br />

standardised test of olfactory function, to patients with ataxias primarily<br />

due to cerebellar pathology (spinocerebellar ataxias and related<br />

disorders) and to patients with Friedreich ataxia, an ataxia associated<br />

mainly with loss of afferent cerebellar pathways. UPSIT scores were<br />

slightly lower in both patient groups than in the control subjects, but no<br />

differences were noted between the scores of the Friedreich and the<br />

other ataxia patients. Within the Friedreich ataxia group, the smell test<br />

scores did not correlate with the number of pathologic GAA repeats (a<br />

marker of genetic severity), disease duration, or categorical ambulatory<br />

ability. UPSIT scores did not correlate with disease duration, although<br />

they correlated marginally with ambulatory status in the patients with<br />

cerebellar pathology. This study suggests that olfactory dysfunction<br />

may be a subtle clinical component of degenerative ataxias, in<br />

concordance with the hypothesis that the cerebellum or its afferents<br />

plays some role in central olfactory processing.<br />

Supported by NIH Research Grants PO1 DC 00161, RO1 DC 04278,<br />

RO1 DC 02974, RO1 AG 17496 (RLD) and a Beeson Scholar Award<br />

(DRL) from AFAR<br />

336 Slide [ ] Human Olfaction: Pathology<br />

DIAGNOSTIC OPTIONS AND LIMITS IN PATIENTS WITH<br />

OLFACTORY DYSFUNCTION AFTER HEAD INJURY<br />

Haxel B.R. 1, Mann W. 1, Mackay-Sim A. 2 1Otorhinolaryngology,<br />

University of Mainz, Mainz, Germany; 2Psychology, Griffith University,<br />

Brisbane, Queensland, Australia<br />

Objective: In this study patients with impaired olfaction after head<br />

trauma were investigated using smell tests and immunohistochemistry<br />

and cell-cultures of biopsies of the olfactory epithelium to trace specific<br />

changes.<br />

Methods: We investigated the sense of smell with Sniffin´ Sticks,<br />

chemosensory evoked potentials and biopsy of the olfactory epithelium<br />

using neurofilament and beta;-tubulin type III to detect immature<br />

neurons on frozen sections and cultures. The results were compared<br />

with age- and gender-matched normosmic controls.<br />

Results: In 5 patients complete anosmia was found in the Sniffin´<br />

Sticks test, but only in 3 patients olfactory evoked potentials were<br />

missing completely. The biopsies showed intact olfactory epithelium<br />

with immature neurons in some cases, cell adhesion was found in 25%<br />

to 100% of the cultures and neurogenesis could be induced under the<br />

influence of growth factors.<br />

Conclusions: Head trauma can cause comprehensible olfactory<br />

impairment; there are different possible sites of a lesion. Olfactometry<br />

shows an intact neural transmission to the brain. Biopsies of the<br />

olfactory epithelium reveal the regeneration capacity of the receptor<br />

neurons.<br />

Acknowledgements: This study was supported by the Deutsche<br />

Forschungsgemeinschaft HA-3447/1-1, Germany and by the Garnet<br />

Passe and Rodney Williams Memorial Foundation, Australia.<br />

88<br />

337 Slide [ ] Human Olfaction: Pathology<br />

OLFACTORY DEFICITS IN SINONASAL DISEASE<br />

Kern R.C. 1, Conley D.B. 1, Robinson A.M. 1 1Otolaryngology-HNS,<br />

Northwestern University, Chicago, IL<br />

Background: Chronic sinonasal disease is a common cause of smell<br />

loss but the pathophysiology is unclear. Human biopsy specimens<br />

indicate direct inflammatory changes in the epithelium, although it<br />

remains unclear how this mediates clinical olfactory deficits. It has been<br />

suggested that inflammatory cytokines may alter secretion and/or affect<br />

neuronal viability. Under normal conditions, olfactory sensory neurons<br />

undergo apoptotic cell death at a baseline rate matched by the<br />

regeneration of mature OSNs from precursors in the epithelium. The<br />

current study examines normal and diseased olfactory mucosa <strong>for</strong><br />

evidence of a disturbance in this balance, which result in a net loss of<br />

OSNs. Study Design: Histologic analysis of human and animal<br />

olfactory tissue. Methods: Normal and inflamed human and animal<br />

olfactory mucosa was assessed <strong>for</strong> immunohistochemical evidence of<br />

apoptosis. Results: Increased activity of the apoptotic effector enzyme<br />

caspase-3 was demonstrated in diseased olfactory mucosa in<br />

comparison with normal controls in both human and animal tissue.<br />

Conclusion: These results support the hypothesis that olfactory deficits<br />

in sinusitis result from increased apoptotic OSN death apparently not<br />

matched by regeneration. Interference with the apoptotic pathway is<br />

currently the subject of intense pharmaco-therapeutic research.<br />

Potential treatment options will be discussed. Supported by the<br />

department of Otolaryngology-HNS.<br />

338 Poster [ ] Human Olfaction: Pathology<br />

CHEMOSENSORY CHANGES FROM EXPOSURE TO<br />

FORMALDEHYDE IN ANATOMY LABS<br />

Dalton P. 1, Gould M. 1, Opiekun R. 1 1Monell Chemical Senses Center,<br />

Philadelphia, PA<br />

More than 10% of the U.S. work<strong>for</strong>ce (~14 million people) has daily<br />

occupational exposure to chemicals or particulates and as many as 40%<br />

of these individuals report symptoms of chronic nasal inflammation or<br />

rhinitis that are known precursors of olfactory loss. Moreover, evidence<br />

suggests that occupational chemical exposure increases the risk of ageassociated<br />

olfactory disorders. Despite this, the evaluation of olfactory<br />

function in susceptible occupational cohorts is rarely per<strong>for</strong>med and<br />

consequently, little is known about the mechanisms underlying<br />

chemical-induced olfactory changes. This study evaluated the<br />

chemosensory impact of repetitive exposure to <strong>for</strong>maldehyde among<br />

veterinary students enrolled in anatomy labs, using a comprehensive<br />

assessment of threshold, secretory, and inflammatory parameters.<br />

Personal exposures <strong>for</strong> each participant were collected using passive<br />

dosimeters in order to evaluate observed changes in olfactory<br />

per<strong>for</strong>mance as a function of exposure concentration. Consistent with<br />

studies in <strong>for</strong>maldehyde-exposed animals, mucociliary clearance time<br />

increased with exposure duration <strong>for</strong> the students, but not unexposed<br />

controls, an outcome which may predispose individuals to both<br />

perceptual changes and nasal health effects following chemical<br />

exposures.<br />

Supported by NIH P50 P50 DC00214


339 Poster [ ] Human Olfaction: Pathology<br />

QUALITATIVE OLFACTORY DYSFUNCTION: FREQUENCY<br />

AND PROGNOSTIC SIGNIFICANCE<br />

Hummel T. 1, Maroldt H. 1, Frasnelli J. 1, Landis B.N. 2, Hüttenbrink K. 3,<br />

Heilmann S. 4 1ORL, University of Dresden Medical School, Dresden,<br />

Germany; 2ORL, University Hospital of Geneva, Geneva, Switzerland;<br />

3ORL, University of Dresden, Germany, Dresden, Germany; 4ORL,<br />

University of Dresden, Dresden, Germany<br />

This study aimed to investigate frequency and prognostic<br />

significance of qualitative olfactory dysfunction (parosmia,<br />

phantosmia). A total of 868 patients were included; 160 of these<br />

patients (18%) complained of parosmic sensations, 59 (7%) mentioned<br />

odor phantoms. In patients without qualitative olfactory dysfunction,<br />

smell loss was most likely due to trauma (30%), infection of the upper<br />

respiratory tract (23%), sinunasal disease (17%) or it was idiopathic<br />

(22%). In patients with parosmia these figures were 7%, 70%, 8%, and<br />

12%; in patients with odor phantoms they were 20%, 54%, 15%, and<br />

17%. Thus, parosmias and phantosmias were most frequently<br />

encountered in olfactory loss following URTI.<br />

Among hyposmic patients those with parosmia exhibited a better<br />

discrimination of odorants but were less proficient in odor identification<br />

when visiting first. However, at second visit there was no significant<br />

difference between patients with (n=67) or without parosmia (n=223).<br />

In patients with parosmia decreased olfactory function was found in<br />

22%, improvement in 39%. Similar figures were seen in patients<br />

without parosmia (21%, and 40%). When visiting first patients with<br />

phantosmia were significantly worse in terms of odor identification<br />

compared to patients without phantosmia. At return visit in 38% of the<br />

phantosmic patients overall olfactory function was decreased;<br />

improvement was found in 41% of the patients. In conclusion,<br />

parosmias are found most frequently in olfactory dysfunction following<br />

URTI. Apparently, the presence of parosmia is not a predictor of the<br />

prognosis of olfactory dysfunction. In contrast, phantosmias appear to<br />

indcate a higher likelihood <strong>for</strong> a further decrease of olfactory fucntion.<br />

89<br />

340 Poster [ ] Human Olfaction: Pathology<br />

HIGH INCIDENCE OF FUNCTIONAL ANOSMIA IN THE<br />

GENERAL POPULATION<br />

Landis B. 1, Konnerth C. 2, Hüttenbrink K. 3, Hummel T. 4 1University<br />

Hospital of Geneva, Geneva, Switzerland; 2ENT, University Hospital of<br />

Dresden, Dresden, Germany; 3Otorhinolaryngology, University of<br />

Dresden Medical School, Dresden, Germany; 4University of Dresden<br />

Medical School, Dresden, Germany<br />

Objective: To assess the incidence of anosmia and hyposmia in the<br />

general population.<br />

Study Design: Open prospective study ruled out on a population<br />

based sample representative <strong>for</strong> the German population.<br />

Methods: A total of 1240 subjects were included. Participants<br />

received an ENT examination and olfactory testing. Patients reporting<br />

sinu-nasal disorders were excluded. Olfactory testing was done by<br />

means of the Sniffin´ Sticks.<br />

Results: In 5% of the subjects nasal polyposis was diagnosed in the<br />

absence of sino-nasal complaints, confirming high incidence of nasal<br />

polyposis previously reported. In the remaining 1182 subjects (603<br />

men, 579 women) functional anosmia was detected at a frequency of<br />

4.7 %, whereas incidence increased with age (Fig. 1). Hyposmia was<br />

found in 16 % of the subjects. Most subjects with functional anosmia<br />

were unaware of their poor per<strong>for</strong>mances; accordingly, they could not<br />

indicate an origin of the anosmia. No significant sex-related difference<br />

in functional anosmia rate was detected. Similar to other authors we<br />

found age to be a main factor leading to functional anosmia. However,<br />

the present data also revealed functional anosmia to occur in up to 5 %<br />

of subjects under 65 years of age.<br />

Conclusion: The current findings suggest that severe olfactory<br />

alteration occurs at a much higher frequency than previously assumed,<br />

especially among younger people. Since olfactory deficits have been<br />

shown to occur in several general pathologies and especially in<br />

neurodegenerative diseases in their early phase, anosmia must be<br />

considered more than just a symptom of chronic sinu-nasal disease and<br />

deserves the general physicians attention. The present data also show<br />

the need of further research into treatments of olfactory disorders.<br />

341 Poster [ ] Human Olfaction: Pathology<br />

ESTROGEN REPLACEMENT THERAPY: DOES IT AFFECT<br />

SMELL FUNCTION IN POST-MENOPAUSAL WOMEN?<br />

Neff J.K. 1, Knipe C. 1, Doty R.L. 1 1Smell and Taste Center, Department<br />

of Otorhinolaryngology, University of Pennsylvania Medical Center,<br />

Philadelphia, PA<br />

Decreased smell function is among the first signs of Alzheimer´s<br />

disease. Although controversial, there is evidence that estrogens<br />

mitigate verbal memory and several other cognitive deficits that are<br />

associated with Alzheimer´s disease. Thus, the question arises as to<br />

whether estrogens mitigate olfactory deficits observed in the older<br />

female population. This retrospective study focuses on estrogen<br />

replacement therapy (ERT) and its possible role in decreasing olfactory<br />

loss in post-menopausal women. Currently, 257 women of an<br />

anticipated 600 women have participated in the study. Preliminary<br />

analyses have focused on olfactory scores of women during the first ten<br />

years after menopause. Within this subset of our sample population,<br />

there is a tendency <strong>for</strong> women who are currently on ERT to have higher<br />

olfactory scores in comparison to women who are not currently taking<br />

ERT. Within this limited age range, no meaningful association between<br />

prior estrogen use and olfactory test scores has yet emerged. Additional<br />

preliminary analyses suggest there may be a correlation between ERT<br />

use and decreased verbal cognitive deficits. Future analyses will<br />

explore the effect of ERT on age-related olfactory and memory deficits<br />

beyond the first ten years after menopause.<br />

This abstract was supported by the following grant from the National<br />

Institutes of Health, Bethesda, MD: RO1 AG 17496-04


342 Slide [ ] Human Olfaction: Pathology<br />

MODELING OF AIRFLOW AND ODORANT DELIVERY<br />

PATTERN IN A PRE- & POST-OPERATIVE NASAL CAVITY:<br />

A QUANTITATIVE EVALUATION OF SURGICAL<br />

INTERVENTION<br />

Zhao K. 1, Scherer P.W. 1, Cowart B.J. 2, Pribitkin E.D. 3, Rosen D. 3,<br />

Dalton P. 2 1Bioengineering, University of Pennsylvania, Philadelphia,<br />

PA; 2Monell Chemical Senses Center, Philadelphia, PA; 3Department<br />

of Otolaryngology-Head and Neck Surgery, Thomas Jefferson<br />

University, Philadelphia, PA<br />

Mechanical obstruction of odorant flow to olfactory receptor sites<br />

due to inflammation or polyps is a primary cause of olfactory loss in<br />

nasal-sinus disease patients. In these cases, surgical intervention (e.g.<br />

removal of polyps) can effectively facilitate recovery of olfactory<br />

ability. Using computational fluid dynamics (CFD) techniques, we are<br />

able to quickly convert nasal C-T scans from an individual patient at a<br />

given time point (e.g., pre-post surgery) into anatomically accurate 3-D<br />

numerical nasal models that can be used to predict nasal airflow and<br />

odorant delivery patterns. Our goal is to correlate the patient´s olfactory<br />

recovery with improvement of odorant delivery rate to receptor sites at<br />

various times during the treatment.<br />

In this preliminary study, we followed the treatment of a patient<br />

who had lost most of her orthonasal and all of her retronasal olfactory<br />

ability, but regained it after surgical treatment of polyps. CFD modeling<br />

of this patient´s nose be<strong>for</strong>e and after surgery showed significant<br />

improvement in ortho & retronasal odorant delivery and suggested that<br />

remodeling the airway was a significant factor leading to the recovery<br />

of olfactory function. In the future, such modeling techniques may<br />

serve as a quantitative evaluation of surgical procedures and an<br />

important pre-surgical guide to the optimization of airflow and odorant<br />

delivery in the human nose.<br />

Supported by NIH P50 DC00214<br />

343 Poster [ ] Human Olfaction: Pathology<br />

DIAGNOSIS AND SURGICAL TREATMENT OF PAROSMIA<br />

Pitovski D.Z. 1, Goins M. 1 1WFU Smell and Taste Center, Dept. of<br />

Otolaryngology, Wake Forest University, Winston-Salem, NC<br />

The purpose of this study was to determine whether transnasal<br />

endoscopic excision of the olfactory mucosa is an effective and safe<br />

treatment in patients diagnosed with unilateral parosmia and to learn<br />

more of its pathogenic features by examining the histological<br />

characteristics of the excised mucosa. The two patients (patient A and<br />

B) who were subjected to the surgery both had complete and permanent<br />

resolution of their parosmia. Ninety-days after surgery, olfactory<br />

assessment revealed no change (in comparison to preoperative status) in<br />

olfactory ability on the operated nostril in patient A, while improved in<br />

olfactory ability in patient B. No changes in olfactory capacity were<br />

noted in the contralateral (unoperated) nostrils. Patient A and patient B<br />

had unremitting parosmia <strong>for</strong> 3 years and 20 years, respectively. The<br />

excised olfactory mucosa showed abnormal histological features that<br />

would suggest some pathological condition in the peripheral olfactory<br />

system. To the best of our knowledge, this is the first report that has<br />

demonstrated effective treatment of patients who suffer from parosmia<br />

through surgical technique. Furthermore, the abnormal histological<br />

features of the excised olfactory mucosa and its excision in the<br />

treatment of the parosmia suggest that there is a peripheral<br />

pathophysiological mechanism.<br />

90<br />

344 Poster [ ] Human Olfaction: Pathology<br />

ODORANT-INDUCED EXACERBATION OF BURNING<br />

MOUTH SYNDROME<br />

Hirsch A.R. 1 1The Smell & Taste Treatment and Research Foundation,<br />

Chicago, IL<br />

BMS presents with symptoms of burning, pain, phantageusia. We<br />

present 3 patients, exposed to odorants precipitated severe burning<br />

exacerbation. (1) 83 year-old woman, 2 years burning, salty<br />

phantageusia, onset immediately following application of nitroglycerin.<br />

Burning worsened with ingesting hot or spicy food, stress, and<br />

lidocaine application. Acute severe exacerbation of BMS precipitated<br />

with exposure to Carbinol in the Olfactory Test of Amoore. (2) 57 yearold<br />

male, 1year sudden onset of idiopathic metallic phantageusia and<br />

BMS. Burning worsened after consuming coffee, chocolate and cola,<br />

reduced with chewing gum or eating salty food. Upon exposure to<br />

aromas of shampoo or soap, the BMS and associated phantageusia were<br />

exacerbated. (3) 59 year-old woman, 1 year of gradual onset<br />

hypogeusia, metallic and spicy phantageusia, salty and spicy dysgeusia,<br />

and BMS. Burning worsened with eating spicy food; reduced with<br />

sucking sweet candy and drinking diet Coke. When presented with<br />

odors of gas or garage, severe burning mouth symptoms recurred.<br />

Possible mechanisms of odorants: stimulate hunger, thus precipitate<br />

salivary flow, change chemical content of the mouth, which<br />

hypersensitive pain fibers interpret as pain; induce eating memories,<br />

stimulate gastroesophageal reflux, change in mouth PH causing pain;<br />

induce cephalopancreatic reflex, causing decreased blood sugar<br />

allowing dysfunctional pain nerve fibers to fire; trigeminal component<br />

of odors may act on tongue to precipitate prolonged, intense burning;<br />

stressors which worsens BMS. BMS may represent a <strong>for</strong>m of<br />

synesthesia, seen in neurological conditions such as odor-sensitive<br />

migraine, or post-amputation phantom limb–in which otherwise<br />

innocuous stimuli precipitate pain.<br />

345 Poster [ ] Olfactory Regeneration<br />

APRIN IN RAT OLFACTORY EPITHELIUM<br />

Weiler E. 1, Farbman A.I. 2 1Neurophysiology, Ruhr-Universität,<br />

Bochum, Germany; 2Neurobiology and Physiology, Northwestern<br />

University, Evanston, IL<br />

Aprin (androgen-induced prostate proliferative shutoff associated<br />

protein AS3) inhibits cell proliferation in the prostate and protects<br />

against carcinogenesis. Although proliferation in the olfactory<br />

epithelium (OE) continues during adulthood OE doesn't develop<br />

tumors. We asked whether aprin is expressed in the rat OE. We report<br />

here the sequence of rat aprin (Accession # AY388627) and the<br />

expression of aprin mRNA in the olfactory epithelium of postnatal rats<br />

using RT-PCR and (competitive) duplex PCR. Semiquantitative<br />

estimates reveal that aprin mRNA expression level in the OE is lower<br />

(~8x) compared to that in testis. Although proliferation dramatically<br />

decreases in rat OE postnatally, the expression level of aprin does not<br />

change much. Besides the OE, aprin mRNA is also expressed in other<br />

neuronal (olfactory bulb, visual cortex, cerebellum, eye, adrenal gland)<br />

and non-neuronal tissues (testis, kidney, heart, lung, and very weak<br />

expression in muscle, intestine, liver) with highest expression in testis.<br />

No differences between males and females were observed in aprin<br />

expression in OE. Aprin function is discussed in relation to cell<br />

turnover and neuronal survival.<br />

Supported by NIH Grants # DC04637, DFG Grant SFB509 TPC4,<br />

FORUM F108/00 M122/13).


346 Poster [ ] Olfactory Regeneration<br />

REDUCED TARGET ABLATION-INDUCED MACROPHAGE<br />

RECRUITMENT AND ACTIVATION IN MIP-1 KNOCK-OUT<br />

(KO) MICE IS RESTORED BY MIP-1 INJECTION<br />

Getchell M.L. 1, Kwong K. 2, Vaishnav R.A. 1, Getchell T.V. 3 1Anatomy<br />

& Neurobiology, Univ. of KY, Lexington, KY; 2Physiology, Univ. of KY,<br />

Lexington, KY; 3Sanders-Brown Ctr. on Aging, Univ. of KY, Lexington,<br />

KY<br />

The chemokine macrophage inflammatory protein (MIP)-1α (CCL3)<br />

recruits macrophages (mφs) to sites of epithelial remodeling. We<br />

showed that MIP-1α mRNA and protein levels in the olfactory<br />

epithelium (OE) increase significantly at 3 d post-olfactory bulbectomy<br />

(OBX). Our specific aim was to investigate the effect of lack of MIP-1α<br />

on mφ infiltration after OBX in MIP-1α KO mice compared to wildtype<br />

(wt) mice and to test if function was restored in KO mice by MIP-<br />

1α injection. OBX was per<strong>for</strong>med on wt and MIP-1α KO mice; all<br />

received 6 injections at 12-h intervals of either 10 µg/ml MIP-1α<br />

protein in carrier (0.5% BSA in sterile saline) or carrier alone <strong>for</strong> 3 d.<br />

Mφ infiltration was evaluated with antibodies to CD68 <strong>for</strong> resident mφs<br />

and F4/80 <strong>for</strong> activated mφs. Compared to wt mice, CD68 + mφ<br />

numbers in the OE were 59% less than in carrier-injected KO mice and<br />

only 28% less in MIP-1α-injected mice (p


348 Poster [ ] Olfactory Regeneration<br />

DIFFERENTIAL RESPONSES TO BULBECTOMY AND<br />

MINOCYCLINE-HCL IN BAX DEFICIENT AND WILD TYPE<br />

MICE<br />

Robinson A.M. 1, Conley D.B. 1, Kern R.C. 1 1Otolaryngology-HNS,<br />

Northwestern University, Chicago, IL<br />

In contrast to wild type mice, bax deficient mice demonstrate an<br />

early but short-lived wave of olfactory sensory neuron (OSN) apoptosis<br />

in response to unilateral bulbectomy that gives way to apoptosis<br />

resistance, at least up to 9 days post-surgery. In order to elucidate early<br />

events in OSN apoptosis following bulbectomy we compared changes<br />

in gene expression and epithelial thickness in wild type and bax<br />

deficient mice and in mice treated with minocycline-HCl. The antibiotic<br />

minocycline-HCl has been shown to have an anti-apoptotic effect in<br />

other systems and is in clinical trials <strong>for</strong> therapeutic use in Parkinson´s<br />

disease. In other systems, the mechanism of minocycline-HCl apoptosis<br />

inhibition is believed to be at the level of inhibition of cytochrome c<br />

release from the mitochondria. This inhibition would suppress<br />

activation of caspase-9 and the downstream caspases, including<br />

caspase-3. Given that bax promotes cytochrome c release, it was<br />

anticipated that apoptosis in the bax knockout mouse may be further<br />

suppressed by minocycline-HCl. By epithelial thickness measurement<br />

there appears to be an initial inhibition of apoptosis in wild type mice<br />

receiving minocycline-HCl. In contrast, the initial apoptosis following<br />

bulbectomy typically observed in bax deficient mice was not<br />

ameliorated. This lack of additional apoptosis inhibition in<br />

minocycline-HCl treated bax deficient mice may be explained by our<br />

results from cDNA expression arrays that focused on apoptotic gene<br />

expression. Bax deficient mice show a dramatic increase in caspase-3<br />

mRNA by 8 hours post-bulbectomy that was absent in wild-type mice.<br />

Supported by the department of Otolaryngology-HNS.<br />

349 Poster [ ] Olfactory Regeneration<br />

OEC DYNAMICS IN THE OLFACTORY SYSTEM OF<br />

METHIMAZOLE-LESIONED & CONTROL MICE<br />

Iwema C.L. 1, Dodds T. 1, Chin J. 1, Greer C.A. 1 1Neurosurgery, Yale<br />

University School of Medicine, New Haven, CT<br />

Olfactory sensory neuron (OSN) axons are held in juxtaposition by<br />

specialized glia, the olfactory ensheathing cells (OECs), as they extend<br />

towards the olfactory bulb (OB). It has been suggested that OECs<br />

provide a favorable substrate <strong>for</strong> axon outgrowth and potentially assist<br />

with targeting and/or glomerular recognition. Damage to the olfactory<br />

system (OS) results in incorrect target recognition by the OSNs; the<br />

effect of injury on the OECs is unknown. One hypothesis is that OECs<br />

fail to correctly ensheath and/or sort OSN axons following lesions. The<br />

effect of growth/guidance molecules associated with OECs on the<br />

olfactory projection after injury remains ambiguous. Despite the<br />

apparent inability of newly developed OSN axons to accurately target<br />

following lesions, they nonetheless do innervate glomeruli in the OB,<br />

albeit in an indiscriminate manner. What happens to the OECs during<br />

this process? We used established immunohistochemical protocols with<br />

antibodies to tyrosine hydroxylase (TH), GAP43, NCAM and S100 to<br />

evaluate changes in the mouse OS during degeneration (5 days, 10d),<br />

reinnervation (30d, 60d), and recovery (90d) following exposure to the<br />

olfactoxin, methimazole. Our data demonstrate that TH expression is<br />

initially decreased in the OB but recovers by 90d post-lesion, whereas<br />

GAP43 expression is markedly upregulated at 10d post-lesion be<strong>for</strong>e<br />

returning to baseline. In addition, we addressed OEC turnover using<br />

both BrdU and a marker <strong>for</strong> apoptosis (NeuroTACS II) and report<br />

evidence of both OEC genesis and death in control tissue. Thus, OECs<br />

appear to be dynamic agents in the normal adult OS. NIH DC00210,<br />

DC06291.<br />

92<br />

350 Poster [ ] Olfactory Regeneration<br />

EARLY OLFACTORY ENRICHMENT DECREASES TUNEL-<br />

POSITIVE CELLS IN OLFACTORY BULBS OF NEONATAL<br />

RATS.<br />

Woo C.C. 1, Hingco E.E. 1, Taylor G.E. 1, Johnson B.A. 1, Leon M. 1<br />

1Neurobiology and Behavior, University of Cali<strong>for</strong>nia, Irvine, Irvine,<br />

CA<br />

A proportion of rat olfactory bulb interneurons normally undergo<br />

postnatal cell death, as is evident by both a decrease in the total<br />

numbers of interneurons after the second postnatal week, and the<br />

presence of TUNEL-positive cells in most layers of the main olfactory<br />

bulb at that time. In the current study, we determined whether increased<br />

olfactory experience could save olfactory bulb cells from an early<br />

death. To that end, we examined the effects of early olfactory<br />

enrichment on cell death in the main olfactory bulb. Rats were exposed<br />

continuously to a battery of natural and artificial odorants that were<br />

changed daily from postnatal days (PNDs) 1-15. On PND 16, TUNELpositive<br />

cells were quantified in both the granule cell layer and the<br />

glomerular layer of the main olfactory bulb. Early olfactory enrichment<br />

resulted in a statistically significant decrease in the number of TUNELpositive<br />

cells present in both of these layers. In addition, we examined<br />

the effects of odor enrichment on the neural response to odorants using<br />

[14C]2-deoxyglucose. Odor enrichment during the first three postnatal<br />

weeks resulted in a statistically significant increase in activity in the<br />

glomerular layer in response to odorants. These data suggest that<br />

olfactory enrichment during the early postnatal period can save<br />

olfactory bulb cells from dying, and can also enhance the glomerular<br />

response to odorants. This research was supported by NIDCD grant<br />

#DC03840 to M.L.<br />

351 Poster [ ] Olfactory Regeneration<br />

CASPASE 8 ACTIVATES ORN APOPTOSIS FOLLOWING<br />

DEAFFERENTATION AND EXCITOTOXIC LESION OF<br />

MOUSE OLFACTORY BULB<br />

Fung F.W. 1, Carson C. 1, Saleh M. 2, Nicholson D. 2, Roskams J. 1<br />

1Zoology, University of British Columbia, Vancouver, British Columbia,<br />

Canada; 2Merck Frosst, Montreal, Quebec, Canada<br />

Our lab has been testing how signals initiated at the olfactory bulb<br />

mitral cell synapse may regulate the apoptosis of olfactory receptor<br />

neurons (ORNs). We have previously shown that, following olfactory<br />

bulbectomy, ORNs undergo apoptosis driven by caspase-9 and 3mediated<br />

retrograde axonal signaling. Because a bulbectomy is also a<br />

distal axotomy, we modified our lesion model to remove only the ORN<br />

target neurons (mitral and tufted cells) in the olfactory bulb using<br />

NMDA-mediated excitotoxicity. We find that NMDA infused directly<br />

into the olfactory bulb effectively kills off the majority of bulbar mitral<br />

and tufted cells within 48 hrs of being administered. Surprisingly, most<br />

OMP-positive mature ORNs do not immediately undergo target<br />

deprivation-induced apoptosis following NMDA infusion, and despite<br />

the loss of functional glomeruli, a significant population of their axons<br />

persist <strong>for</strong> up to 8 days in the nerve fiber layer. A delayed <strong>for</strong>m of<br />

apoptosis does appear to eventually occur (by 4 days after loss of<br />

postsynaptic target) in a sub-population of vulnerable ORNs in NMDAlesioned<br />

mice. We have identified caspase 8 as the apical caspase first<br />

activated at the glomerular synapse. We demonstrate (by coimmunoprecipitation)<br />

that Dynactin, a DED (Death Effector Domain) -<br />

containing protein, is a retrograde motor complex protein that can<br />

regulate the transport and/or activation of caspase 8 in ORNs following<br />

target lesion. Finally, by using taxol to inhibit axonal microtubule<br />

motors following bulbectomy, we can delay caspase-mediated<br />

retrograde apoptosis in ORNs.<br />

Supported by NIH (NIDCD) 5RO1 DC04579-03 to JR


352 Slide [ ] Odorant Receptors<br />

COMPARATIVE GENOMICS OF OLFACTORY RECEPTOR<br />

GENE CLUSTERS<br />

Young J.M. 1, Newman T. 1, Schlador M. 1, Linardopoulou E. 1, Walker<br />

M. 1, Hsu J. 1, Williams E. 1, Trask B.J. 1 1Division of Human Biology,<br />

Fred Hutchinson Cancer Research Center, Seattle, WA<br />

Comparison of the nearly complete genomes of rat, mouse, and<br />

human reveals extensive lineage-specific change in the olfactory<br />

receptor (OR) gene repertoires of these species. The primary<br />

contributor to repertoire change has been recent duplications that create<br />

new genes. Genes have also been lost through deletion and mutation.<br />

Only ~400 human genes appear to be functional, compared to ~1210 <strong>for</strong><br />

mouse and ~1390 <strong>for</strong> rat. Our cDNA screen of olfactory<br />

neuroepithelium confirmed expression there <strong>for</strong> over one third of the<br />

intact mouse OR genes and also revealed extensive alternative splicing<br />

within the 5´UTRs of OR transcripts. We observe up to a 300-fold<br />

difference in mRNA levels among mouse OR genes, which appears to<br />

be due to differences in both the number of expressing cells and the<br />

number of transcripts per cell. The family expanded primarily through<br />

local duplications in all three species, but inter-chromosomal<br />

duplications also occurred in both the human and rat lineages.<br />

Intriguingly, these inter-chromosomal duplications involved large<br />

genomic segments (up to 800 kb) containing primarily OR<br />

pseudogenes. The complex structures and evolutionary relationships<br />

among these regions suggest that they are subject to ongoing gene<br />

conversion-like ectopic exchange subsequent to the original<br />

duplication. The propensity of these segmental duplications to undergo<br />

recombination can also lead to gross allelic variation among normal<br />

individuals and /or the <strong>for</strong>mation of chromosome abnormalities<br />

associated with disease. Funded by NIH DC004209 and GM057070.<br />

353 Poster [ ] Odorant Receptors<br />

HANDS OFF MY ENDANGERED SPECIES: LOW ALLELIC<br />

VARIATION OF SEA TURTLE OR GENES SUPPORTS<br />

IMPORTANCE OF OLFACTORY SENSE.<br />

Vieyra M. 1, Vogt R.G. 1 1Biological <strong>Sciences</strong>, University of South<br />

Carolina, Columbia, SC<br />

How to determine the importance of a given sense to the life history<br />

of an endangered species when government regulations prohibit direct<br />

physiological analysis? Estimate the evolutionary selective pressure<br />

acting on the sense. Clone an "important" gene and examine its allelic<br />

variation within a population; low variation suggesting selection is<br />

strong and that the sense is important to the life history of the animal,<br />

high variation suggesting selection is low and that the sense is less<br />

important. Cloning and characterizing genes of endangered species is<br />

feasible because archives of DNA samples OR genes were cloned from<br />

3 sea turtles (loggerhead, green, leatherback), 4 freshwater or terrestrial<br />

turtles (musk, painted, box, gopher tortoise) and American alligator.<br />

These genes are more similar to the mammalian than to the fish OR<br />

gene subfamily. Sea turtles have a reduced number of OR genes and a<br />

greater number of OR pseudogenes (internal stop codons) than<br />

freshwater or terrestrial turtles. Population analysis of 5 sea turtle genes<br />

indicates a strong dominance of a single allele <strong>for</strong> conserved genes, and<br />

a surprisingly low level of allelic variation among pseudogenes. These<br />

findings suggest olfaction is important and that pseudoization may be a<br />

relatively recent event if not an ongoing process. Turtles have similar<br />

sized OR and VNO systems; VNO neurons are thought to convey<br />

aquatic signals. The presence of strongly selected OR genes suggests<br />

that airborne odors may be important signals <strong>for</strong> sea turtles. Funding:<br />

NOAA<br />

93<br />

354 Poster [ ] Odorant Receptors<br />

EXPRESSION OF CANDIDATE GUSTATORY RECEPTOR<br />

GENES IN ANOPHELES GAMBIAE<br />

Kent L.B. 1, Robertson H.M. 1 1Entomology, University of Illinois at<br />

Urbana-Champaign, Urbana, IL<br />

The ability to detect and discriminate between chemical cues in the<br />

environment is key to successful host finding in mosquitoes. The recent<br />

availability of the genome sequence of Anopheles gambiae, the<br />

principle vector of human malaria in Africa, has allowed <strong>for</strong> the<br />

identification and annotation of the genes believed to encode odorant<br />

(AgOr) and gustatory (AgGr) receptor genes in the mosquito. These<br />

genes are members of the G-coupled protein receptor (GCPR)<br />

superfamily of chemoreceptors, characterized in part by their seventransmembrane<br />

domains. Phylogenetic analysis reveals that only a few<br />

orthologous pairs of the genes have been conserved between<br />

Drosophila and Anopheles. Otherwise, most Ors and Grs <strong>for</strong>m speciesspecific<br />

gene subfamilies. We have begun studies of the expression<br />

patterns of the putative AgGr genes, based upon an RT-PCR assay. In<br />

this way, we can examine the conservation of expression of Grs in A.<br />

gambiae over the roughly 250 million years since it diverged from the<br />

lineage containing Drosophila melanogaster. Patterns of similarity<br />

across the two species are apparent. Three of the genes (AgGr22,<br />

AgGr23, AgGr24) with high sequence similarity to two genes in<br />

Drosophila melanogaster (DmGr21a and DmGr63a) are expressed in<br />

antennae and maxillary palps, which matches the expression patterns<br />

seen in Drosophila. Interestingly, DmGr21a has recently been<br />

identified by Dr. Marien de Bruyne, at the Freie Universität Berlin, as a<br />

candidate carbon dioxide receptor in Drosophila. We are in the process<br />

of determining whether the closely related anopheline genes share the<br />

same function. This research is funded by NIH RO1AI056081.<br />

355 Poster [ ] Odorant Receptors<br />

EXPRESSION OF AN ANOPHELES GAMBIAE CANDIDATE<br />

ODORANT RECEPTOR IN A SUBSET OF DISTINCT<br />

SENSILLA ON THE PROBOSCIS INDICATES A POTENTIAL<br />

OLFACTORY FUNCTION<br />

Pitts J. 1, Rützler M. 1, Zwiebel L. 1 1Biological <strong>Sciences</strong> and Center <strong>for</strong><br />

Molecular Neuroscience, Vanderbilt University, Nashville, TN<br />

The biting behavior of Anopheles gambiae (An. gambiae) is largely<br />

influenced by olfactory cues emanating from host animals. The strong<br />

preference of the An. gambiae s.s. species <strong>for</strong> human hosts contributes<br />

significantly to the transmission of human malaria in sub-Sahara Africa.<br />

The family of odorant receptor genes in An. gambiae encodes G<br />

protein-coupled receptors <strong>for</strong> which some ligands have been identified.<br />

One of these genes, GPRor7, is highly conserved with respect to single<br />

genes from many insect orders. Immunocytochemistry demonstrates<br />

that its protein product is localized to most sensilla of olfactory organs<br />

of An. gambiae. By RT-PCR analysis, GPRor7 is also expressed in the<br />

proboscis of An. gambiae, a known gustatory organ in other<br />

mosquitoes. Immunocytochemisry indicates that GPRor7 is found<br />

within each of a small subset of sensilla on the labellar lobes, a pair of<br />

bulbous organs at tip of the proboscis. Scanning electron micrographic<br />

analysis reveals that the GPRor7 expressing sensilla are short, grooved<br />

hairs residing in a socket on the cuticle surface and that there are about<br />

25 of these hairs per labellar lobe. Their morphology is reminiscent of<br />

other grooved olfactory sensilla, and the presence of an odorant<br />

receptor within them indicates either a potential olfactory function <strong>for</strong><br />

them, or a potential gustatory function <strong>for</strong> GPRor7. This work was<br />

supported by grants from the National Institutes of Health: the NIDCD<br />

and the NIAID.


356 Poster [ ] Odorant Receptors<br />

OLFACTORY CODING IN PERIPHERAL ORGANS OF<br />

ANOPHELES GAMBIAE<br />

Kwon H. 1, Zwiebel L.J. 1 1Department of Biologial <strong>Sciences</strong> and Center<br />

<strong>for</strong> Molucualr Neuroscience, Vanderbilt University, Nashville, TN<br />

A principal goal in neuroscience is to understand how sensory<br />

in<strong>for</strong>mation is processed and integrated to control a suitable behavioral<br />

output in a nervous system of an animal. Of these, olfactory cues are<br />

primarily used in an insect to find a host, mate, and oviposition site.<br />

There<strong>for</strong>e, it is important to understand how olfactory coding occurs a<br />

peripheral olfactory organ as well as in central nervous system. In<br />

insects, olfactory transduction is initiated by G protein-coupled<br />

receptors in the cell membrane, which have been characterized in<br />

different insect species so far. In this study, using an important malaria<br />

vector mosquito, A. gambiae, from which 79 different G proteincoupled<br />

receptor genes have been found, we focus on functional<br />

analysis of olfactory in<strong>for</strong>mation processing in antenna as well as<br />

proboscis where an odorant receptor gene (AgOR7) is also expressed.<br />

With analysis of gene expression in these appendages, here we aim to<br />

characterize central processing and olfactory coding in a primary<br />

olfactory and gustatory organ, antenna and proboscis, using<br />

electrophysiological methods and backfilling techniques. Recording<br />

from epithelia of antenna and proboscis was employed with several key<br />

mosquito odorant components. Here we report that proboscis shows<br />

olfactory responses to an odorant chemical. This result supports the idea<br />

that proboscis contributes olfactory perception in accordance with gene<br />

expression of the odorant receptor in proboscis. More detailed evidence<br />

of neural anatomy and single sensillum recording will be presented.<br />

(Supported by NIH grants A1056402 & DC04692)<br />

357 Poster [ ] Odorant Receptors<br />

MOLECULAR ANALYSIS OF DROSOPHILA ODORANT<br />

RECEPTORS<br />

Benton R. 1, Vosshall L.B. 1 1Rockefeller University, New York, NY<br />

The Drosophila olfactory system is a powerful model to determine<br />

how neuronal circuits represent the external world in the brain. The<br />

peripheral circuitry has a striking stereotypical spatial organization.<br />

Olfactory sensory neurons (OSNs) express only one of ~60 odorant<br />

receptors (ORs) along with Or83b, a broadly expressed member of the<br />

OR gene family. The axons of OSNs expressing the same OR converge<br />

on specific glomeruli in the antennal lobe. Each class of OSN also<br />

displays distinct spiking properties in response to odor stimulation,<br />

suggesting that both temporal and spatial patterns of neuronal activity<br />

contribute to the representation of odor identity.<br />

ORs are polytopic membrane proteins that lie at the interface<br />

between the odorous environment and neuronal activity patterns. Little<br />

is known about their ligand specificity, what second messenger cascade<br />

couples receptor activation to neuronal spiking, or the mechanisms by<br />

which their activity and localization are regulated. Genetic analysis of<br />

candidate downstream signaling components has failed to reveal the<br />

involvement of a single canonical pathway, suggesting that multiple<br />

and/or novel cascades are involved. ORs are not expressed on the<br />

surface of heterologous cells, indicating that OSNs possess unique<br />

trafficking pathways integral to the localization and function of ORs.<br />

To dissect OR function and regulation, we have initiated a<br />

combination of molecular and transgenic approaches. We are using the<br />

yeast two-hybrid system to identify cytoplasmic factors that associate<br />

with their intracellular loops, and have generated a panel of epitopetagged<br />

ORs <strong>for</strong> biochemical analysis in vivo.<br />

Funding: EMBO, NIH and NSF<br />

94<br />

358 Poster [ ] Odorant Receptors<br />

THE SPERM "NOSE": KEY ROLE OF PARTICULATE<br />

ADENYLATE CYCLASE<br />

Schwane K. 1, Spehr M. 1, Riffell J. 2, Barbour J. 1, Zimmer R. 2, Neuhaus<br />

E.M. 1, Hatt H. 1 1Cell Physiology, Ruhr-University Bochum, Bochum,<br />

Germany; 2Biology, University of Cali<strong>for</strong>nia, Los Angeles, Los Angeles,<br />

CA<br />

Besides their 'conventional' role in olfaction, odorant receptors have<br />

long been suggested to function in mammalian sperm physiology and<br />

fertilization. Recently, Spehr et al. (2003) identified and characterized a<br />

human testicular odorant receptor, hOR17-4, that is activated by certain<br />

floral odorants (e.g. bourgeonal, scent of lilies of the valley) In<br />

additional behavioral bioassays, bourgeonal was found to act as a strong<br />

chemoattractant to navigating human sperm. Nevertheless, the<br />

molecular mechanisms that link hOR17-4 activation to sperm responses<br />

remain obscure.<br />

In imaging experiments as well as behavioral assays, we show here<br />

that a membrane-bound AC (mAC) couples OR activation to changes in<br />

sperm swimming behaviour, such as chemotaxis, chemokinesis, and<br />

hyperactive flagellar beating.<br />

Introducing a new approach of protein identification in mature sperm<br />

by mass spectrometry (mudpit) we provide evidence <strong>for</strong> expression and<br />

participation of specific receptors, G-proteins, and mACs in the<br />

underlying signaling cascade. Spatial distribution patterns of the<br />

identified signaling components largely correspond to the<br />

spatiotemporal character of odorant-induced Ca2+ changes viewed via<br />

single cell high-resolution imaging techniques.<br />

Taken together our data show that mAC activation is linked to sperm<br />

chemotaxis and hyperactivity and provide a basis <strong>for</strong> further<br />

investigations in the field of fertility and sterility.<br />

This work was generously sponsored by the Heinrich and Alma<br />

Vogelsang Foundation (to K.S.).<br />

359 Poster [ ] Odorant Receptors<br />

CHEMICAL COMMUNICATION AND THE LANGUAGE<br />

SPOKEN BY SPERM AND EGGS<br />

Zimmer R. 1, Riffell J. 1, Krug P. 2 1Ecology and Evolution, University of<br />

Cali<strong>for</strong>nia, Los Angeles, Los Angeles, CA; 2Biological <strong>Sciences</strong>,<br />

Cali<strong>for</strong>nia State University, Los Angeles, Los Angeles, CA<br />

Chemical communication between sperm and egg is a critical factor<br />

mediating sexual reproduction. Sperm attractants may be significant<br />

evolutionarily <strong>for</strong> maintaining species barriers, and important<br />

ecologically <strong>for</strong> increasing gamete encounters. Still unresolved,<br />

however, are the functional consequences of these dissolved signal<br />

molecules. Here, we provide the first experimental evidence that sperm<br />

chemoattraction directly affects the magnitude of fertilization success.<br />

The recent discovery of L-tryptophan as a potent attractant to red<br />

abalone (Haliotis rufescens) sperm offered the opportunity to quantify<br />

how navigation affects gamete interactions. Sperm behavioral responses<br />

to manipulations of the natural tryptophan gradient around individual<br />

eggs revealed that both chemotaxis and chemokinesis significantly<br />

promote contacts. Our results showed further that attractant release via<br />

diffusion effectively doubles the target size of red abalone eggs, which<br />

in turn significantly increases fertilization success. Although long<br />

theorized as potential barriers to hybridization, species-specific sperm<br />

attractants in red and green (H. fulgens) abalone are only minor<br />

contributors to maintaining reproductive isolation. Because abalone<br />

typically live in dense, multi-species aggregations, chemically mediated<br />

navigation would prevent sperm from pointlessly tracking<br />

heterospecific eggs. Thus, even though reproductive isolation<br />

fundamentally resides at the level of membrane recognition proteins,<br />

species-specific sperm attractants may have evolved to locate the right<br />

target within mixed gamete suspensions of closely related species.


360 Poster [ ] Odorant Receptors<br />

MOUSE TESTICULAR OLFACTORY RECEPTORS:<br />

EXPRESSION PATTERN, ODORANT RESPONSIVENESS, AND<br />

REGULATION OF SPERM MOTILITY.<br />

Fukuda N. 1, Yomogida K. 2, Okabe M. 3, Kataoka H. 1, Touhara K. 1<br />

1Department of Integrated Biosciences, University of Tokyo, Kashiwa,<br />

Chiba, Japan; 2Research Institute <strong>for</strong> Microbial Diseases, Osaka<br />

University, Suita, Osaka, Japan; 3Genome In<strong>for</strong>mation Research<br />

Center, Osaka University, Suita, Osaka, Japan<br />

Although a subset of olfactory receptor (OR) gene family is<br />

expressed in mammalian testis, developmental expression pattern and<br />

physiological function of testicular ORs have not been fully<br />

characterized. To clarify testicular OR function in mice, we first<br />

analyzed expression pattern of mouse OR genes in testis and found that<br />

ORs were expressed stage-specifically in round spermatids during<br />

spermatogenesis. We next conducted functional analysis of MOR23, a<br />

mouse testicular OR, that had been shown to recognize a floral odorant,<br />

lyral, in the olfactory system. Lyral elicited Ca2+ -increases in a fraction<br />

of spermatids and mature sperm in a dose-dependent manner.<br />

Comparison of lyral-responsiveness of germ cells derived from<br />

transgenic mice expressing MOR23 with that of wild type mice<br />

provided evidence that MOR23 mediated lyral-induced Ca2+ -increases<br />

in spermatids and spermatozoa. Finally, lyral induced sperm<br />

accumulation in chemotaxis assay, suggesting that MOR23 functioned<br />

as a chemosensor in mature sperm, and that the internal Ca2+ -increase<br />

via MOR23 affected the sperm motility. The MOR23-expressing<br />

transgenic mouse line generated in this study will be a powerful tool to<br />

identify endogenous ligand(s) that may be structurally related to lyral,<br />

with implications <strong>for</strong> physiological roles of mouse testicular OR(s).<br />

361 Slide [ ] Odorant Receptors<br />

DISCOVERY OF ACETALS, ALCOHOLS, AND ESTERS AS<br />

ISOVALERIC ACID ODOR BLOCKERS<br />

Qi M. 1, Rogers D.H. 1, Warren C.B. 2, Darmohusodo V. 1 1Senomyx, Inc.,<br />

La Jolla, CA; 2Chemosensory Perception Lab, UC San Diego, La Jolla,<br />

CA<br />

Isovaleric acid (IVA) is probably a potent agonist <strong>for</strong> one or more of<br />

the approximately 350 human olfactory receptors (Zozulya 2001). To<br />

find an odor blocker, we hypothesized a simple binding pocket based<br />

on known GPCR receptors and then designed antagonist candidates that<br />

would fit into this binding pocket. Each candidate was tested <strong>for</strong> its<br />

ability to reduce the perceived odor intensity of a 56 ppm solution of<br />

IVA in water (Target). The evaluation of blocking utilized a cross<br />

adaptation protocol in which the subject a) sniffed the target solution<br />

with both nostrils and scored its odor intensity on a Labeled Magnitude<br />

Scale (LMS); b) waited one minute, c) sniffed a 1% v/v solution of the<br />

candidate in diethyl phthalate twice with each nostril, and d) sniffed the<br />

target again with both nostrils. The reduction in IVA intensity between<br />

the first and third sniffing was used to calculate a % malodor reduction<br />

(MOR) <strong>for</strong> each candidate. The compounds that worked best have a<br />

hydrocarbon tail similar to that of IVA and a polar head. Of the<br />

homologous series tested, the acetals worked best with 2-isobutyl-<br />

[1,3]dioxane giving a MOR score of 97%. For the alcohols, 2methylcyclopropane<br />

methanol provided the best blocking with a MOR<br />

of 85%. For the ester series, the ethyl and methyl esters of IVA gave<br />

MOR scores of 94 and 92% respectively. The magnitude of blocking<br />

was much higher than the 25 to 30% odor reduction generally reported<br />

in the cross adaptation literature. These results suggest that we have<br />

identified antagonists <strong>for</strong> the primary IVA receptor.<br />

95<br />

362 Slide [ ] Odorant Receptors<br />

HELA CELLS DESIGNED FOR FUNCTIONAL GENOMICS OF<br />

ODORANT RECEPTORS AND PHEROMONE RECEPTORS<br />

Shirokova E. 1, Schmiedeberg K. 1, Bedner P. 2, Niessen H. 2, Willecke<br />

K. 2, Harteneck C. 3, Raguse J. 4, Krautwurst D. 1 1Molecular Genetics,<br />

German Institute of Human Nutrition Potsdam-Rehbrücke, Nuthetal,<br />

Germany; 2Institute of Genetics, Bonn University, Bonn, Germany;<br />

3Institute of Pharmacology, Free University Berlin, Charité Campus<br />

Benjamin Franklin, Berlin, Germany; 4Clinic and Policlinic <strong>for</strong> Oral<br />

and Maxillofacial Surgery and Plastic Surgery, Charité, Berlin,<br />

Germany<br />

Odorant receptors (OR) of the main olfactory epithelium, and<br />

pheromone receptors (VR) of the vomeronasal organ, are the largest<br />

group of orphan G-protein-coupling receptors (GPCR), with a small<br />

number of ligands identified out of thousands of odorants and hundreds<br />

of pheromones. De-orphanization of OR was hampered by their<br />

combinatorial odorant coding, suboptimal plasma membrane<br />

expression, and lack of olfactory-specific signal transduction in<br />

recombinant cell systems. Heterologous functional expression of VR<br />

has not been reported, so far. We have achieved stable reconstitution of<br />

OR-specific signaling in HeLa cells, via G-protein αolf and adenylyl<br />

cyclase type-III to the olfactory cyclic nucleotide-gated CNGA2<br />

channel. In these cells, receptors of the V1R-type inhibit, via G protein<br />

αi, the cAMP pathway. In another cell line, we established stable<br />

expression of a TRP channel that can be activated by pheromones via<br />

V1R and phospholipase C-β2. CNGA2 or TRP channels translate<br />

changes in intracellular cyclic nucleotides into changes in Ca 2+ -influx<br />

that can be monitored by means of fluorescence imaging multiwellplate<br />

reader techniques. This allows us to functionally characterize the<br />

de-orphanized OR and V1R by EC 50 -ranking odorant profiles.


363 Poster [ ] Odorant Receptors<br />

MECHANISM FOR OLFACTORY RECEPTOR-ODORANT<br />

INTERACTIONS<br />

Crasto C.J. 1, Lai P.C. 2, Singer M. 3, Shepherd G.M. 1 1Neurobiology,<br />

Yale University, New Haven, CT; 2Molecular and Cell Biology,<br />

University of Connecticut, Storrs, CT; 3Deaprtment of Opthalmology,<br />

Massachusetts General Hospital, Harvard University, Boston, MA<br />

Objective of the Study<br />

To study the interactions between olfactory receptors (OR) and odor<br />

molecules through molecular dynamics simulations.<br />

Methods<br />

Ten odorant ligands, eight carbon atom chain aldehydes with variable<br />

branches, side chains and unsaturation, were docked into the binding<br />

region of rat olfactory receptor I7, the first olfactory receptor identified<br />

by Buck & Axel in1991 and modeled by Singer (2000). The<br />

experimental binding of the ligands with I7 has been previously<br />

reported by Araneda and co-workers. (2000). We used the Singer OR<br />

model, and the lowest energy configuration <strong>for</strong> each docked ligand.<br />

Each system was first minimized, and molecular dynamics simulations<br />

were carried out <strong>for</strong> 500 picoseconds.<br />

Results<br />

Our results <strong>for</strong> in vacuo molecular dynamics simulations indicate that<br />

the ligand attempts periodically to exit and re-enter the OR binding<br />

region. There is an exit channel in I7 irrespective of whether the loops<br />

of the helical domains are included in the model. The exit events were<br />

strongly correlated with significant structural changes within the<br />

binding pocket.<br />

Conclusions<br />

The exit-reentry events may be related to the binding strengths of the<br />

ligands with the OR. This previously unreported exit-reentry behavior<br />

in OR-odor dynamics is an important consideration in the study of the<br />

mechanism of olfaction at the molecular level.<br />

Acknowledgments<br />

This work was supported by the Human Brain Project and the<br />

National Library of Medicine.<br />

364 Poster [ ] Odorant Receptors<br />

CAN TWO NOSTRIL SNIFFING HELP ELECTRONIC NOSES?<br />

Johnson B.N. 1, Khan R.M. 1, Sobel N. 1 1Bioengineering, University of<br />

Cali<strong>for</strong>nia, Berkeley, Berkeley, CA<br />

There is worldwide interest in development of devices to detect and<br />

identify airborne chemical compounds. Currently, techniques such as<br />

gas-chromatography and mass-spectroscopy are expensive and difficult<br />

to deploy in a portable package. Portable sensing devices have a wide<br />

range of applications such as monitoring of food and drug processing,<br />

manufacturing, dangerous substance identification, land mind detection,<br />

and medical diagnostics. However, chemical sensor technology is new<br />

and applications are often limited by sensor sensitivity, selectivity,<br />

discrimination, drift, and durability. A new theme in electronic nose<br />

research is to use the mammalian olfactory system as a model to<br />

address these issues. This direction is focusing on using multiple<br />

sensors and sophisticated multi-variant statistical models (PCA,<br />

learning machines, etc). However, there has been little attention to<br />

sampling technique. Often, sampling is done by flowing the carrier gas<br />

over the sensor or exposing the sensor to an aqueous solution<br />

containing the odorants. These methods are the easiest to per<strong>for</strong>m, but<br />

may not be the most effective. In the mammalian olfactory system,<br />

there are two separate inputs--the left and the right nostrils. Each nostril<br />

receives slightly different flow rates and this results in odorants having<br />

different solubility characteristics between the two nostrils. In essence,<br />

the olfactory system receives two offset samples in a single sniff. Here<br />

we explore the possible benefits of using mammal-like sampling<br />

strategies <strong>for</strong> electronic nose technologies.<br />

96<br />

365 Poster [ ] Odorant Receptors<br />

CHARACTERIZATION OF THE MECHANISM OF ODOR<br />

SENSING IN NOVEL DNA-BASED FLUORESCENT SENSORS.<br />

Williams L.B. 1, Kauer J.S. 1, White J.E. 1 1Neuroscience, Tufts<br />

University, Boston, MA<br />

Our laboratory has developed an artificial nose that exploits in<br />

principle, some 22 attributes of the biological olfactory system. One<br />

important feature is the use of broadly tuned sensor arrays to achieve<br />

odor detection. In the first such experiments of which we are aware, we<br />

show that 20-30 base single-stranded DNA-Cy3 (ssDNA) conjugates<br />

can respond to odorant molecules in our artificial nose. DNA-Cy3<br />

conjugates have the combinatorial potential to provide large arrays of<br />

novel sensors. The mechanism by which these sensors operate is as yet<br />

unknown. Preliminary work has produced a set of 9 DNA-Cy3 sensors.<br />

Using this sensor set, the following odorant response observations have<br />

been made: 1) Sensor responses return to baseline when no longer<br />

exposed to odorant, suggesting that the interactions between the odorant<br />

and the sensor are rapidly reversible and there<strong>for</strong>e non-covalent. 2)<br />

Change in fluorescence intensity (∆F) can be either positive or<br />

negative, suggesting that a simple quenching process is unlikely to be<br />

the mechanism. 3) Sensors tested so far (dried from water based<br />

buffers) do not respond as well to non-polar odorants as to polar<br />

odorants, and 4) Sensors do not function well at below 15% humidity<br />

suggesting a role <strong>for</strong> residual water (solvent). 5) Odor detection<br />

increases in a relatively linear manner with increasing concentration of<br />

odor. Based on these data, we believe that the ∆F seen in the presence<br />

of odors may be due to odorant molecules dissolving into the hydration<br />

layer surrounding the DNA and exerting solvent effects on Cy3. These<br />

are modulated by DNA sequence mechanisms under investigation that<br />

involve tertiary structure.<br />

Supported by grants from NIDCD and NSF.<br />

366 Slide [ ] Taste Genetics and Physiology<br />

ASSOCIATIONS BETWEEN PTC/PROP GENE, 6-N-<br />

PROPYLTHIOURACIL (PROP) BITTERNESS AND ALCOHOL<br />

INTAKE<br />

Duffy V.B. 1, Davidson A. 2, Kidd J. 2, Kidd K. 3, Speed W. 2, Pakstis A. 2,<br />

Reed D. 4, Snyder D. 5, Bartoshuk L. 5 1Dietetics Program, University of<br />

Connecticut, Storrs, CT; 2Genetics, Yale University, New Haven, CT;<br />

3Psychiatry, Yale University, New Haven, CT; 4Monell Chemical Senses<br />

Center, Philadelphia, PA; 5Surgery, Yale University, New Haven, CT<br />

Nontasters (PROP is weakly bitter) report less negative (bitterness,<br />

irritation) and more positive (sweet) sensations from alcoholic<br />

beverages than do supertasters (PROP is strongly bitter). Younger (23<br />

F, 26 M) and middle-aged (30 F, 5 M) adults rated bitterness of 5 PROP<br />

concentrations (0.032-3.2 mM) on the general Labeled Magnitude<br />

Scale. Alcoholic beverage intake was assessed by frequency survey;<br />

adults stated drinking at least once per year. DNA was analyzed <strong>for</strong> the<br />

PTC/PROP gene (TAS2R38) that codes <strong>for</strong> molecularly different<br />

receptors depending on specific amino acids at three positions in the<br />

protein. The common <strong>for</strong>ms are AVI and PAV [Proline, Alanine,<br />

Valine, Isoleucine] (Kim et al, 2003). Via repeated measures analysis of<br />

variance (ANOVA), PROP bitterness varied significantly across<br />

genotype groups [AVI (n=26) less than PAV/AVI (n=38) less than<br />

PAV (n=22)] but not enough to explain supertasting. Age group<br />

(younger>middle-age) and 3.2 mM PROP bitterness (greater bitterness,<br />

less intake) contributed significantly to predicting alcohol intake via<br />

multiple regression analyses. With ANOVA, alcohol intake varied<br />

significantly across genotype groups (AVI>PAV/AVI>PAV) yet<br />

significance was primarily in younger adults. These data support taste<br />

genetic effects on alcohol intake. Phenotypical and genetic markers of<br />

taste may be necessary to study orosensory effects on diet across aging.<br />

(NRICGP/USDA 2002-00788, NIH DC00283, GM 57672)


367 Slide [ ] Taste Genetics and Physiology<br />

GENETICS OF PTC TASTE SENSITIVITY IN HUMANS<br />

Drayna D. 1, Kim U. 2, Wooding S. 3, Jorde L. 4, Floriano W. 5, Goddard<br />

W.A. 5 1National Institute on Deafness and Other Communication<br />

Disor, National Institutes of Health, Rockville, MD; 2NIDCD, National<br />

Institutes of Health, Rockville, MD; 3Genetics, University of Utah, Salt<br />

Lake City, UT; 4Human Genetics, University of Utah, Salt Lake City,<br />

UT; 5Chemistry, Cali<strong>for</strong>nia Institute of Technology, Pasadena, CA<br />

The inability of some individuals to taste phenylthiocarbamide (PTC)<br />

was discovered more than 70 years ago and since that time it has been<br />

the subject of detailed studies in genetics, anthropology, and sensory<br />

physiology. We have identified the major gene underlying this trait as<br />

TAS2R38, a G protein coupled bitter taste receptor located on<br />

chromosome 7q. The taster and non-taster <strong>for</strong>ms of this protein differ at<br />

3 amino acid positions, and we have identified other alleles of this gene<br />

that encode various combinations of these 3 variant amino acids, at<br />

least some of which produce intermediate PTC sensitivity. We have<br />

used population genetic methods to study the paradoxical high<br />

frequency of the non-taster <strong>for</strong>m of this gene. Our analyses indicate that<br />

both the major taster and major non-taster alleles have been maintained<br />

by balancing natural selection. Since bitter taste is thought to protect<br />

individuals from ingestion of toxic substances in our diet, this raises the<br />

question of what selective benefit is provided by the non-taster allele.<br />

We hypothesize that the non-taster <strong>for</strong>m of the protein serves as a<br />

functional receptor <strong>for</strong> a different toxic bitter substance not yet<br />

identified. We have also employed molecular structure prediction<br />

techniques to determine the 3D structure of the taster and non-taster<br />

<strong>for</strong>ms of this receptor, along with PTC ligand binding sites. Our results<br />

indicate that PTC binds to both <strong>for</strong>ms of the receptor with equal<br />

affinity, and that the non-taster <strong>for</strong>m of the protein does not signal due<br />

to a failure of G protein activation.<br />

Supported by NIDCD/NIH Z01-000046 (to D.D.), by NIH grants<br />

ES12125 (to S.W.), GM59290 (to L.J.), and NSF grant BCS0218370<br />

(to L.J.)<br />

97<br />

368 Slide [ ] Taste Genetics and Physiology<br />

GENETIC CONTROL OF LICK RATE IN MICE<br />

Boughter J. 1, St. John S. 2, Williams R.W. 3, Lu L. 1 1Anatomy &<br />

Neurobiology, University of Tennessee, Memphis, TN; 2Psychology,<br />

Reed College, Portland, OR; 3Anatomy and Neurobiology, University of<br />

Tennessee, Memphis, Memphis, TN<br />

Licking is a highly stereotyped behavior that is thought to be<br />

determined in part by a central pattern generator (CPG). Lick rate has<br />

been shown to differ among inbred strains of mice, and may play a role<br />

in apparent differences in gustatory sensitivity. We used a 20 minuteaccess<br />

procedure to quantify lick rate in several strains of waterdeprived<br />

mice. Inbred strains could be classified as fast, intermediate,<br />

or slow lickers based on the median value of the inter-lick interval (ILI)<br />

distribution. With virtually no overlap among individual distributions,<br />

C57BL/6J (B6; n = 36) mice were classified as slow lickers (average<br />

median ILI = 122.8 ms) whereas DBA/2J (D2; n = 27) mice were fast<br />

lickers (average median ILI = 101.1). Mice from an F1 generation (n =<br />

14) possessed an intermediate phenotype (112.8 ms). Fast / slow lick<br />

rate phenotypes were static over an extended time period, and were also<br />

evident in non-deprived tests with sucrose. Further testing of an F2<br />

generation (n = 62) resulted in an estimate of broad-sense heritability =<br />

0.54, with no fewer than 2 polymorphic genes influencing the ILI<br />

phenotype. Initial QTL mapping analysis using a panel of BXD RI<br />

strains indicates that at least two and possibly as many as 4 QTL with<br />

comparatively large effects on the ILI are segregating in this cross.<br />

These data will soon be complemented with a genome scan of a panel<br />

of ~ 150 F2 mice that are now being typed using a panel of ~70<br />

microsatellite markers. Additionally, we show how differences in lick<br />

rate between B6 and D2 strains may exert a non-gustatory influence in<br />

lick ratio measurements of sucrose sensitivity. Supported by<br />

DC004935(JDB).<br />

369 Slide [ ] Taste Genetics and Physiology<br />

INTAKE OF SWEET AND BITTER SOLUTIONS: VARIATION<br />

IN INBRED STRAINS OF GOLDEN HAMSTERS<br />

Frank M.E. 1, Wada Y. 2, Makino J. 2, Mizutani M. 2, Umezawa H. 3,<br />

Katsuie Y. 3, Hettinger T.P. 1, Blizard D.A. 4 1Oral Diagnosis,<br />

Neurosciences, UCONN Health Center, Farmington, CT; 2Institute of<br />

Psychology, University of Tsukuba, Tsukuba, Ibaraki, Japan;<br />

3Laboratory Animal Research Station, Nippon Institute <strong>for</strong> Biological<br />

Science, Kobuchizawa, Yamanashi, Japan; 4Pennsylvania State<br />

University, University Park, PA<br />

Intake of sweet and bitter solutions by 7 inbred strains of golden<br />

hamsters (Mesocricetus auratus), a principal species <strong>for</strong> studies of<br />

mammalian gustatory systems, was measured. Two concentrations of<br />

sucrose, maltose, D-Phe and Na saccharin, which are sweet; and<br />

quinine·HCl, L-Phe, caffeine and sucrose octaacetate (SOA), which are<br />

bitter to humans, were tested. Difference scores, solution intake minus<br />

mean baseline water intake in mL (DIF), were evaluated by analysis of<br />

variance (α = .05). Compared to ACN, CN, APA, APG and CBN (5<br />

strains with similar DIF <strong>for</strong> all tested solutions), the strains ACNT and<br />

GN (an ACNT ancestral strain) preferred sucrose, caffeine and SOA<br />

more strongly; ACNT also preferred saccharin and maltose more<br />

strongly and rejected quinine more strongly. There were no strain<br />

differences in DIF <strong>for</strong> D-Phe or L-Phe. Narrow sense heritabilities <strong>for</strong><br />

the 6 compounds <strong>for</strong> which strain differences were revealed ranged<br />

from 0.31 to 0.57. Genetic correlations indicated the strain variations in<br />

intake of sucrose, saccharin, SOA and caffeine were coupled,<br />

suggesting an association with several possible interpretations. The<br />

genetic differences that influence taste behaviors in existing strains of<br />

hamsters may help identify relevant genes. [Supported by NIH grant<br />

R01 DC04099 (MEF) and a Japan Society Senior Fellowship (DAB)]


370 Slide [ ] Taste Genetics and Physiology<br />

RESPONSES TO ETHANOL IN WILD TYPE (WT) AND -<br />

GUSTDUCIN KNOCKOUT (KO) MICE<br />

Danilova V. 1, Danilov Y. 1, Damak S. 2, Margolskee R. 2, Hellekant G. 1<br />

1Animal Health and Biomedical <strong>Sciences</strong>, University of Wisconsin-<br />

Madison, Madison, WI; 2Physiology and Biophysics, Mount Sinai<br />

School of Medicine, New York, NY<br />

Although it has been shown that C57Bl/J mice prefer ethanol, it is<br />

not known what kind of taste nerve responses ethanol elicits. We have<br />

shown in primates that ethanol stimulates chorda tympani (CT) and<br />

glossopharyngeal (NG) taste fibers and the response depends on type of<br />

taste fibers. Our purpose was to study taste responses to ethanol in<br />

mice. We also investigated if α-gustducin is involved in transduction of<br />

ethanol taste.<br />

Methods. CT and NG responses and results of two bottle preference<br />

tests (TBP) with ethanol (0.5-5 M) were recorded in WT and KO mice.<br />

Results. 1. In WT only high concentrations of ethanol elicited CT<br />

responses. This contrasts NG effects in which low concentrations<br />

elicited significant responses. 2. In both nerves the responses developed<br />

slowly. 3. In KO ethanol did not produce CT responses at any<br />

concentration. The NG responses did not differ between the two groups.<br />

4. In contrast to WT, KO did not prefer ethanol at any concentration<br />

and rejected higher concentrations in TBP tests.<br />

Conclusion. Ethanol is not an effective stimulus in WT mice in<br />

contrast to primates. Absence of preference and CT responses to<br />

ethanol together with previously demonstrated decrease of responses to<br />

sweeteners in KO mice suggest that ethanol has a sweet taste<br />

component, which may be the cause <strong>for</strong> its consumption by WT. Thus<br />

α-gustducin is important in transduction of ethanol taste and<br />

consumption. Rejection of ethanol by the KO might be explained by the<br />

significant responses in the NG, which may originate from non-taste<br />

fibers or non-sweet fibers.<br />

Supported by NIH DC 03155 and NIH DC005336<br />

371 Slide [ ] Pheromones<br />

A DROSOPHILA ODORANT-BINDING PROTEIN MEDIATES<br />

RESPONSES TO A PHEROMONE<br />

Smith D. 1, Xu P. 1, Atkinson R. 2, Jones D. 3 1Pharmacology, University<br />

of Texas Southwestern Medical Center at Dallas, Dallas, TX;<br />

2Phamacology, University of Texas Southwestern Medical Center at<br />

Dallas, Dallas, TX; 3Pharmacology, University of Colorado Health<br />

<strong>Sciences</strong> Center, Denver, CO<br />

Pheromones are chemosensory cues released by animals to influence<br />

the behavior of other members of the same species. In Drosophila, a<br />

male-specific lipid, 11-cis vaccenyl acetate (VA), mediates olfactorybased<br />

social aggregation. OBPs are a large family of proteins found in<br />

all terrestrial species and have been suggested to transport odorants.<br />

Here we show that the odorant binding protein 76a (OBP76a), an<br />

extracellular protein secreted into the fluid bathing a subset of olfactory<br />

neurons, is required <strong>for</strong> behavioral attraction to VA. This phenotype is<br />

due to a complete loss of VA-evoked activity in pheromone sensitive<br />

neurons. These defects are reversed by germline trans<strong>for</strong>mation with a<br />

cloned, wildtype copy of obp76a or, importantly, by introducing<br />

recombinant OBP76a protein into mutant sensilla. Remarkably,<br />

spontaneous activity of the pheromone-sensitive neurons is reduced<br />

over 400-fold in the absence of the binding protein. This implicates the<br />

binding protein as a direct activator of the pheromone-sensitive<br />

neurons. These studies directly link odorant binding protein expression<br />

with activation of a specific subset of olfactory neurons and<br />

pheromone-induced behavior. We suggest OBP76a is an extracellular<br />

modulator of neuronal activity that translates the presence of<br />

pheromone to neuronal activation.<br />

98<br />

372 Slide [ ] Pheromones<br />

PHEROMONE REGULATION OF A TRANSCRIPTION<br />

FACTOR IN THE HONEY BEE BRAIN<br />

Grozinger C.M. 1, Robinson G.E. 2 1Department of Entomology,<br />

Program in Neuroscience, University of Illinois at Urbana-Champaign,<br />

Urbana, IL; 2Department of Entomology; Program in Neuroscience,<br />

University of Illinois, Urbana-Champaign, Urbana, IL<br />

Christina M. Grozinger*‡ and Gene E. Robinson‡<br />

*Beckman Institute <strong>for</strong> Advanced Science and Technology<br />

‡Department of Entomology<br />

‡Program in Neuroscience<br />

University of Illinois, Urbana-Champaign<br />

Honey bee social organization is predominantly regulated by<br />

chemical communication. Using honey bee cDNA microarrays, we<br />

have begun to analyze the molecular mechanisms of pheromonal<br />

regulation of honey bee behavior by using queen mandibular<br />

pheromone (QMP), one of the primary pheromonal regulators of<br />

physiology and behavior of worker bees. One of the genes QMP<br />

regulates in the bee brain is the transcription factor, Kr-h1. Kr-h1 is<br />

expressed primarily in the mushroom bodies, the region of the insect<br />

brain associated with multimodal integration and learning, and its<br />

expression is regulated by QMP especially strongly in this region.<br />

Furthermore, Kr-h1 expression is correlated with <strong>for</strong>aging in bees, and<br />

seems to be activated prior to the transition to the <strong>for</strong>aging behavioral<br />

state. Previous work has demonstrated that honey bee mushroom<br />

bodies undergo a period of synaptic remodeling and expansion prior to<br />

the initiation of <strong>for</strong>aging, and synaptic remodeling continues as bees<br />

gain <strong>for</strong>aging experience. The specific localization and timing of Kr-h1<br />

expression suggests that it may be involved in this process.<br />

Supported by a Beckman Institute Postdoctoral Fellowship to CMG,<br />

and grants from the Burroughs Welcome Trust, NIH, and USDA to<br />

GER


373 Slide [ ] Pheromones<br />

CHEMICAL COMMUNICATION IN ZEBRAFISH: HOW<br />

PHEROMONES AFFECT FEMALE MATE CHOICE<br />

Gerlach G. 1 1Marine Resources Center, Marine Biological Laboratory,<br />

Woods Hole, MA<br />

In contrast to males, female fitness can decrease significantly by<br />

inappropriate matings. In species in which females usually receive no<br />

resources from a male besides his sperm, females might choose mating<br />

partners whose genes will confer greater fitness on her offspring. While<br />

several studies have shown how pheromones trigger and synchronize<br />

mating behavior in fish very little is known if pheromones can be used<br />

as signals to assess mate quality. In odor choice tests we analyzed the<br />

olfactory preference of female zebrafish <strong>for</strong> a) males of different<br />

relatedness and b) males of different social rank.<br />

A single female zebrafish was exposed to stimulus water of different<br />

males on either side of a flume. The time a female spent on either side<br />

was determined. As odor stimuli holding water of males were used that<br />

were either related (brother) or unrelated to the female. In the second<br />

experiment two male zebrafish were isolated in a 9 l tank and observed<br />

<strong>for</strong> agonistic interactions. After two days the dominant and subordinate<br />

male were separated in single tanks and their holding water was used as<br />

stimuli.<br />

The results can be summarized as follows: 1) Adult females preferred<br />

the odor of unrelated males. 2) Females preferred odor cues of<br />

dominant over subordinate males.<br />

We conclude that kin recognition mechanisms such as phenotype<br />

matching occur in zebrafish that enables females to avoid inbreeding. In<br />

addition, dominant males seem to release a specific odor cue that made<br />

them distinguishable from subordinate males.<br />

Our successful breeding of a zebrafish mutant without a nose and an<br />

olfactory bulbus provides a unique opportunity <strong>for</strong> further analyses of<br />

the effect of pheromones on mate choice.<br />

374 Slide [ ] Olfactory Development, Disease, and Plasticity<br />

IT CAME FROM THE SEA – OLFACTORY ADAPTATIONS<br />

FOR A TERRESTRIAL LIFE IN THE ROBBER CRAB (BIRGUS<br />

LATRO)<br />

Stensmyr M.C. 1, Erland S. 2, Greenaway P. 3, Hansson B.S. 1 1Swedish<br />

University of Agricultural <strong>Sciences</strong>, Alnarp, Sweden; 2Ecology, Lund<br />

University, Lund, Sweden; 3Biological <strong>Sciences</strong>, University of New<br />

South Wales, Sydney, New South Wales, Australia<br />

An important step in the evolution of the olfactory sense has been the<br />

transition from sea to land. Although terrestrial and aqueous olfactory<br />

systems share many characteristics, marked differences exist. These<br />

distinctions likely reflect the differences of the ligands that the systems<br />

need to detect, air borne, mostly hydrophobic volatiles on land and<br />

water-soluble molecules in the sea. The olfactory system of land crabs,<br />

whose terrestrial existence is a comparatively recent evolutionary<br />

development, represents an excellent opportunity to investigate the<br />

effects of the sea to land transition. Have land crabs come to the same<br />

solutions as other terrestrial animals, or is their olfactory sense<br />

characterized by unique innovations? Here we show that the terrestrial<br />

robber crab (Birgus latro) have evolved an olfactory sense displaying a<br />

high degree of resemblance to the insect system. The similarities extend<br />

to physiological, behavioural and morphological characters. The insect<br />

nose of the robber crab is a striking example of convergent evolution,<br />

and nicely illustrates how similar requirements result in similar endproducts.<br />

99<br />

375 Slide [ ] Olfactory Development, Disease, and Plasticity<br />

EXAMINATION OF CIRCADIAN RHYTHMS IN THE<br />

ANTENNA OF THE MOTH MANDUCA SEXTA<br />

Stengl M. 1, Flecke C. 2, Schuckel J. 2, Siwicki K.K. 3 1University of<br />

Marburg, Marburg, Germany; 2Biology, Philipps-University of<br />

Marburg, Marburg, Germany; 3Biology, Swarthmore College,<br />

Swarthmore, PA<br />

A coupled network of circadian pacemakers orchestrates<br />

physiological processes in a 24 h rhythm. In Drosophila melanogaster<br />

circadian pacemakers in the antenna contain the circadian protein<br />

PERIOD (PER) and appear to drive 24 h rhythms of olfactory<br />

sensitivity. In the hawkmoth Manduca sexta a circadian rhythm of<br />

octopamine was measured in the hemolymph and octopaminergic<br />

neurons project into the antenna. Because octopamine is known to<br />

sensitize pheromone-perception in different moths apparently via<br />

cAMP-dependent mechanisms we wanted to know whether this also<br />

applies to the night-active M. sexta. In addition, we tested whether also<br />

in M. sexta olfactory receptor neurons (ORNs) are PERimmunoreactive.<br />

In extracellular tip-recordings from single sex-pheromone-dependent<br />

ORNs responses to the pheromone-component bombykal were tested<br />

under constant light at two different time points with or without the<br />

presence of 8bromo-cAMP in the recording electrode. During control<br />

recordings pheromone-dependent action potential responses showed a<br />

decline from ZT 1-4, but not at ZT 8-11. This decline was not seen in<br />

the presence of 8bromo-cAMP. Thus, possibly circadian adaptation of<br />

the action potential generator occurs during the day which is<br />

counteracted by octopamine-dependent cAMP-rises at night. Antibodies<br />

against the circadian clock protein PER stained ORNs and other<br />

antennal cells. Future studies examine whether PER-antibodies delete<br />

the circadian decline in the sensitivity of ORNs.[Supported by DFG<br />

grant STE531/13-1 to M.S.]


376 Slide [ ] Olfactory Development, Disease, and Plasticity<br />

LIFE STAGE AND ODORANT-INDUCED CHANGES IN<br />

OLFACTORY SENSITIVITY IN COHO SALMON,<br />

ONCORHYNCHUS KISUTCH<br />

Dittman A. 1, May D. 2, Baldwin D. 3, Scholz N. 3 1Northwest Fisheries<br />

Science Center; NOAA-Fisheries, Seattle, WA; 2School of Aquatic and<br />

Fishery Science, University of Washington, Seattle, WA; 3Northwest<br />

Fisheries Science Center, Seattle, WA<br />

Over the lifetime of an organism, the sensitivity of the olfactory<br />

system to specific odors may change in response to developmental<br />

changes, hormones, environmental stimuli, and odorant exposure.<br />

Salmon provide an excellent model <strong>for</strong> studying such changes because<br />

almost every aspect of their lives is influenced by olfaction and they<br />

experience dramatic developmental and environmental transitions<br />

(smolting, maturation, freshwater vs. oceanic) as part of their<br />

migrations from freshwater to oceanic feeding grounds and back.<br />

Furthermore, these homing migrations are governed by olfactory<br />

discrimination of home stream odors that juvenile salmon learn (imprint<br />

to) prior to their seaward migrations. Our previous studies demonstrated<br />

that salmon imprinted to the odorant phenylethyl alcohol (PEA)<br />

developed a long-term sensitization of peripheral olfactory neurons to<br />

this odorant. To further examine this imprinting phenomenon, we<br />

exposed juvenile coho salmon, Oncorhynchus kisutch to three distinct<br />

classes of odorants (amino acids, bile acids and PEA) during smolting,<br />

the presumptive sensitive period <strong>for</strong> imprinting. To assess life stage and<br />

olfactory imprinting associated changes in the sensitivity of the<br />

olfactory system, we recorded electrical field potentials (electroolfactograms)<br />

generated in response to these three classes of odorants<br />

and ovarian fluid at four distinct life stages: oceanic juvenile, maturing<br />

adult , immature adult, and mature. Our results suggest that olfactory<br />

sensitivity changes over the lifetime of the salmon and previous odor<br />

exposure can influence olfactory responses.<br />

Funded by BPA Grant #199305600<br />

377 Slide [ ] Olfactory Development, Disease, and Plasticity<br />

DEFECTIVE OLFACTORY DEVELOPMENT IN 3GNT1 NULL<br />

MICE<br />

Schwarting G. 1, Raitcheva D. 2, Hennet T. 3, Henion T. 1 1University of<br />

Massachusetts Medical School (Worcester), Waltham, MA; 2Shriver<br />

Center, Waltham, MA; 3University of Zurich, Zurich, Switzerland<br />

Subsets of olfactory neurons in mice express unique lactosamine<br />

containing glycans (LCGs) that reacts with the monoclonal antibody,<br />

1B2. We have previously shown that LCGs are expressed by neurons in<br />

all zones of the embryonic OE but project axons preferably to the<br />

ventral OB, suggesting that this glycan may participate in axon<br />

guidance mechanisms. The gene encoding an enzyme critical <strong>for</strong><br />

synthesis of LCGs was recently isolated. In situ hybridization reveals<br />

that this enzyme, β1-3-N-acetylglucosaminyltransferase-1 (β3GnT-1) is<br />

expressed by neurons in the OE beginning at very early embryonic<br />

stages. We show here that mice deficient in β3Gnt-1 fail to express<br />

LCGs during embryonic and early postnatal olfactory development.<br />

β3Gnt-1 null mice have severe defects in <strong>for</strong>mation of connections<br />

between the OE and OB at birth. Between P1 and P10, many OMP+<br />

glomeruli are absent from the glomerular layer of the OB of β3GnT-1<br />

null mice. This defect in glomerulogenesis is accompanied by increased<br />

neuronal cell death in P1 mice followed by an increase in neurogenesis<br />

at P10. In addition, the expression of some odorant receptors, such as<br />

P2, are significantly down-regulated in β3Gnt-1 null mice. Although<br />

OBs of β3GnT-1 null mice are about 25% smaller than control<br />

littermates, overall structure and layering of these OBs is relatively<br />

normal. In summary, mice that do not express LCGs in the embryonic<br />

and early postnatal OE fail to <strong>for</strong>m normal axonal connections with the<br />

OB, suggesting that LCGs play an important role in axon growth and<br />

guidance in the developing mouse olfactory system. Supported by<br />

DC00953.<br />

100<br />

378 Slide [ ] Olfactory Development, Disease, and Plasticity<br />

EFFECT OF AIR POLLUTION ON OLFACTORY FUNCTION<br />

IN RESIDENTS OF MEXICO CITY<br />

Hudson R. 1, Arriola A. 1, Martínez-Gómez M. 2, Distel H. 3 1Inst. Biomed.<br />

Res., Univ. Nacional Autónoma de México, Mexico City, Mexico;<br />

2Centro Tlaxcala Biol. Conducta, Univ. Autónoma de Tlaxcala,<br />

Tlaxcala, Mexico; 3Inst. Med. Psychol., Univ. München, München,<br />

Germany<br />

To our knowledge there has been no study of the effect of everyday<br />

air pollution on olfactory function. It was there<strong>for</strong>e the aim of this study<br />

to compare the olfactory per<strong>for</strong>mance of long-term residents of Mexico<br />

City (MC) - an environment with high air pollution, with the olfactory<br />

per<strong>for</strong>mance of residents of the Mexican state of Tlaxcala (Tx) - a<br />

region culturally and geographically similar to MC but with low air<br />

pollution. Healthy volunteers (MC n = 82, Tx n = 86) from 20-63 years<br />

of age and balanced <strong>for</strong> gender were tested <strong>for</strong> the perception of the<br />

odors of everyday beverages presented in squeeze bottles. When tested<br />

with ascending concentrations of stimuli in a 3-way oddball paradigm,<br />

residents of Tx detected the odor of an orange juice preparation (Clight)<br />

and of Nescafe at significantly lower concentrations than residents of<br />

MC. They could also attribute a quality to and then finally correctly<br />

identify the stimuli at lower concentrations. However, differences<br />

between the groups decreased across the three tasks, suggesting the<br />

increasing participation of central, cognitive processes unimpaired by<br />

pollution. Residents of Tx also per<strong>for</strong>med significantly better in<br />

discriminating between the two similarly smelling Mexican beverages<br />

of horchata and atole in oddball tests. Significant differences between<br />

the two populations were apparent even in the youngest subjects. No<br />

significant differences were found between sexes. Thus, air pollution in<br />

Mexico City appears to have a substantial impact on peripheral<br />

olfactory function, even in young adults.<br />

379 Slide [ ] Olfactory Development, Disease, and Plasticity<br />

OLFACTORY DYSFUNCTION OCCURS IN TRANSGENIC<br />

MICE OVEREXPRESSING HUMAN TAU PROTEIN<br />

Doty R.L. 1, Macknin J. 2, Kerr K. 3, Higuchi M. 4, Lee V. 5, Trojanowski<br />

J. 5 1Smell and Taste Center, Department of Otorhinolaryngology: Head<br />

& Neck Surgery, University of Pennsylvania, Philadelphia, PA; 2Smell<br />

& Taste Center, Department of Otorhinolaryngology: Head & Neck<br />

Surgery, University of Pennsylvania, Philadelphia, PA; 3Smell & Taste<br />

Center, Department of Otorhinolaryngology: Head and Neck Surgery,<br />

University of Pennsylvania, Philadelphia, PA; 4Laboratory <strong>for</strong><br />

Proteolytic Neuroscience, Riken Brain Science Institute, Saitama,<br />

Japan; 5Center <strong>for</strong> Neurodegenerative Disease Research, University of<br />

Pennsylvania, Philadelphia, PA<br />

Disorders of olfaction are among the first clinical signs of<br />

neurodegenerative diseases such as Alzheimer´s disease (AD) and<br />

idiopathic Parkinson´s disease (PD). In this study, we evaluated the<br />

olfactory function of Ta1-3RT transgenic mice that overexpress tau, a<br />

key pathogenic protein in AD, and compared such function to that of<br />

wild type controls who do not express this protein. The Ta1-3RT mice,<br />

but not the controls, exhibited responses indicative of decreased<br />

olfactory function. These data (a) lend support to the notion that tau<br />

may be involved in the pathogenesis of the olfactory dysfunction of<br />

some neurodegenerative diseases and (b) demonstrate, <strong>for</strong> the first time,<br />

that olfactory function is present in a transgenic mouse model of<br />

neurodegenerative tauopathies including the filamentous tau tangles<br />

seen in AD. Future studies need to similarly assess other pathogenic<br />

markers, as well as their distribution within various sectors of the brain,<br />

to determine the specificity of this phenomenon.<br />

Supported, in part, by the following grants from the National<br />

Institure of Health, Bethesda, MD: RO1 DC 04278, RO1 DC 02974,<br />

RO1 AG 17496 and PO1 AG 11542.


380 Slide [ ] Olfactory Development, Disease, and Plasticity<br />

PREDATOR AND NON-PREDATOR ODORS: SIMILARITIES<br />

IN SPECTRAL AND BEHAVIORAL PATTERNS<br />

Lowry C.A. 1, Kay L.M. 2 1Neurobiology, University of Chicago,<br />

Chicago, IL; 2Psychology, University of Chicago, Chicago, IL<br />

Previous research has suggested that olfactory structures respond to<br />

predator odors with bursts of oscillatory activity in the beta frequency<br />

band (15-40Hz). However, those same oscillation bursts were seen in<br />

response to some non-predator odors. We compare physiological and<br />

behavioral responses to six monomolecular odorants (toluene, amyl<br />

acetate, benzaldehyde, acetone, indole, vanillin) with responses to two<br />

predator odors (trimethyl thiazoline [TMT] and fox urine). Olfactory<br />

bulb (OB) local field potential (LFP) responses were recorded during<br />

five consecutive 2-minute presentations of odorants in a closed<br />

chamber, and behavior was recorded by webcam. Theta (3-15 Hz), beta<br />

(15-40Hz), low gamma (35-60Hz), and high gamma (60-115Hz)<br />

oscillation patterns are examined in response to each odor. We find<br />

similarities between predator and non-predator odors across frequency<br />

bands, as well as some differences between the two predator odors<br />

tested. Behavioral responses, such as freezing and attentive sniffing, are<br />

strongly correlated with concurrent oscillatory activity. Our results<br />

suggest that variations in oscillatory patterns can be attributed primarily<br />

to behavioral variations. We also find that isoamyl acetate (IAA)<br />

produces frequency responses different from all the other<br />

monomolecular odors tested. This suggests that IAA, which is used as a<br />

control odor in many olfactory studies, may be inappropriate <strong>for</strong> this<br />

purpose.<br />

Support: Brain Research Foundation Fay/Frank Seed Grant<br />

381 Poster [ ] Cell Biology of the Olfactory Epithelium<br />

TYROSINE HYDROXYLASE PROMOTER-DRIVEN<br />

REPORTER GENE EXPRESSION IN OLFACTORY<br />

EPITHELIUM OF TRANSGENIC MICE<br />

Sasaki H. 1, Berlin R. 1, Baker H. 1 1Burke Med. Res. Inst., Weill Med.<br />

Coll., Cornell Univ., White Plains, NY<br />

Transgenic mice expressing either green fluorescent protein (GFP) or<br />

lacZ reporter genes driven by 9kb of tyrosine hydroxylase (TH)<br />

promoter have been extensively used in characterizing the CNS<br />

dopaminergic phenotype. To investigate whether TH promoter driven<br />

reporter gene expression occurred in the olfactory epithelium (OE),<br />

TH/GFP and TH/lacZ transgenic mouse strains were examined using<br />

immunohistochemical or X-gal histochemical methods. The antibodies<br />

used were chicken anti-GFP, rabbit anti-TH, rabbit anti-βgal and goat<br />

anti-olfactory marker protein (OMP). Both GFP-immunoreactive (IR)<br />

and βgal-IR cells localized to the superficial OE. The cells showed a<br />

sparse distribution that was similar to the zone 1 pattern of olfactory<br />

receptor neurons (ORN). Morphologically, the cells resembled ORNs<br />

with a dendritic and an axonal process directed towards the nasal cavity<br />

and the lamina propria, respectively. Axonal processes did not enter the<br />

OB. Double label confocal analysis showed that OMP, a marker of<br />

mature ORNs, was not coexpressed with either GFP or βgal. X-gal<br />

staining confirmed the cellular morphology and distribution of the<br />

transgene expressing cells in the OE. The vomeronasal organ contained<br />

a few βgal-IR cells. Reporter gene expression was negatively correlated<br />

with age with cell number highest at postnatal day 2 (P2), the first age<br />

examined, and declining until few cells could be detected at P22. These<br />

studies suggest that some cells in the OE transiently exhibit low level<br />

expression of dopaminergic properties during early postnatal<br />

development. (Supported by grant #AG09686)<br />

101<br />

382 Poster [ ] Cell Biology of the Olfactory Epithelium<br />

THE FINE STRUCTURAL DISTRIBUTION G-PROTEIN<br />

RECEPTOR KINASE 3 (GRK3), &BETA-ARRESTIN-2,<br />

CA++/CALMODULIN-DEPENDENT PROTEIN KINASE II<br />

(CAMKII), AND PHOSPHODIESTERASE PDE1C2 IN<br />

OLFACTORY EPITHELIA<br />

Menco B. 1 1Neurobiology and Physiology, Northwestern University,<br />

Evanston, IL<br />

The sequentially activated molecules of olfactory signaling onset are<br />

mostly concentrated in the long thin distal parts of olfactory epithelial<br />

(OE) receptor cell (ORC) cilia (Chem Senses 22:295,1997). Is this also<br />

true <strong>for</strong> the sites of signal termination? GRK3 is thought to act at the<br />

level of the odor receptors, whereas &beta-arrestin-2, CAMKII, and<br />

PDE1C2 are thought to act at the level of adenylyl cyclase. We used<br />

four antibodies to GRK3, two to &beta-arrestin-2 (one, courtesy Dr.<br />

Lefkowitz, Duke Univers.), five to CAMKII (one to both the &alpha<br />

and &beta <strong>for</strong>m, and two each specific to &alpha- and &beta-<br />

CAMKII), and two to PDE1C2 (courtesy Dr. Beavo, Univers. of<br />

Washington). Earlier, light microscopic, studies showed that antibodies<br />

to all of these molecules labeled the OE luminal border that includes<br />

ORC cilia (Science, 259:825,1993; J Biol Chem 41:25425,1997;<br />

Neuron 21:495,1998; Proc Natl Acad. Sci USA 94:3388,1997).<br />

However, the current, fine structural, data indicate that none of these<br />

antibodies labeled ORC cilia exclusively, and included other ORC<br />

structures, such as dendritic endings, as well. Significantly, the<br />

antibodies also labeled apices and microvilli of neighboring OE<br />

supporting cells and some bound to luminal regions, including cilia, of<br />

the respiratory epithelium adjacent to the OE. Thus, among others, the<br />

observations suggest that neuronal ORCs and neighboring OE<br />

supporting cells share at least some molecules of signal transduction. In<br />

the supporting cells, these molecules occur in addition to those of<br />

biotrans<strong>for</strong>mation and transepithelial transport. Supported by NSF<br />

Grant IBN-0094709.<br />

383 Poster [ ] Cell Biology of the Olfactory Epithelium<br />

IMMUNOCYTOCHEMICAL LOCALIZATION OF 11 BETA-<br />

HYDROXYSTEROID DEHYDROGENASE IN THE<br />

MAMMALIAN OLFACTORY MUCOSA<br />

Foster J.D. 1, Sullivan D. 1 1Anatomy, Western University of Health<br />

<strong>Sciences</strong>, Pomona, CA<br />

Mineralocorticoid (type I) receptors have been identified in the<br />

olfactory mucosa by a number of approaches. In immunohistochemical<br />

studies, mineralocorticoid receptor immunoreactivity was localized to<br />

the supranuclear region of sustentacular cells, as well as the acinar cells<br />

of the Bowman's glands. Functioning of the mineralocorticoid (type I)<br />

receptor is enhanced by the action of the enzyme 11 betahydroxysteroid<br />

dehydrogenase type II (11 beta HSD2). 11-beta HSD2<br />

catalyses the inactivation of glucocorticoids, thus playing a major role<br />

in the protection of the mineralocorticoid (type I) receptor. In previous<br />

studies using histochemical and biochemical approaches, 11 beta HSD<br />

was identified in the olfactory mucosa. The objective of this current<br />

study was to identify the corticosteroid inactivating (type II) <strong>for</strong>m of the<br />

11 beta HSD by use of a monoclonal antibody specific to 11-beta<br />

HSD2. When we incubated sections from the olfactory mucosa with<br />

this antibody, 11 beta HSD2 immunoreactivity was found associated<br />

with the acinar cells of Bowman´s glands and the supranuclear region<br />

of sustentacular cells. This distribution corresponds with the location of<br />

mineralocorticoid (type I) receptors reported in earlier studies and the<br />

histochemical location of 11 beta HSD. This study provides further<br />

evidence that mineralocorticoid hormones may be involved in the<br />

modulation of olfactory function. Supported by Western University<br />

College of Osteopathic Medicine of the Pacific Summer Research<br />

Program


384 Poster [ ] Cell Biology of the Olfactory Epithelium<br />

LOCALIZATION OF RETINOIC ACID RECEPTORS IN<br />

MOUSE AND HUMAN NASAL EPITHELIUM<br />

Yee K.K. 1, Hahn C. 2, Rawson N.E. 1 1Monell Chemical Senses Center,<br />

Philadelphia, PA; 2Dept. Psychiatry, University of Pennsylvania,<br />

Philadelphia, PA<br />

All-trans retinoic acid (ATRA), a metabolite of vitamin A, binds to<br />

retinoic acid receptors (RARs) to mediate gene-transcription in target<br />

cells. We previously found that an ATRA supplement enhanced<br />

olfactory recovery rate in adult mice after olfactory bulb (OB)<br />

deafferentation. In this study, we examined the localization of<br />

RAR&alpha, RAR&beta, and RAR&gamma after injury and ATRA<br />

treatment using immunocytochemistry. Mice received a left olfactory<br />

nerve transection (LNX) with the right side serving as control. One day<br />

after LNX, the mice were given either ATRA mixed with sesame oil or<br />

just sesame oil. In the control OB, RAR&alpha immunoreactivity (ir)<br />

was observed in periglomerular and granule cells, while we did not<br />

detect immunostaining <strong>for</strong> RAR&beta or RAR&gamma. In the oiltreated<br />

right olfactory epithelium (OE), RAR&alpha -ir was found in<br />

NST- and SUS4-negative microvillar-like cells located in the<br />

supporting cell layer, in cells in the lamina propria, and in some<br />

respiratory cells. RAR&beta -ir was localized only in the respiratory<br />

cells while no RAR&gamma -ir was observed in the OE. Surprisingly,<br />

the density of RAR&alpha -ir microvillar-like cells was higher in the<br />

transected OE and highest in transected OE with ATRA treatment,<br />

suggesting these cells are non-neural. We also examined RARs-ir in<br />

human nasal tissue and found a similar cellular localization. These<br />

findings suggest that microvillar cells, a population about which<br />

comparatively little is known, are targets <strong>for</strong> ATRA modulation of gene<br />

expression in the OE and may play a role in OE recovery following<br />

injury. Supported by NIH DC04645 and DC00214.<br />

385 Poster [ ] Cell Biology of the Olfactory Epithelium<br />

IMMUNOLOCALIZATION OF BEX PROTEINS IN THE<br />

MOUSE BRAIN: COLOCALIZATION WITH OMP<br />

Koo J. 1, Manda S. 1, Margolis F.L. 1 1Medicine, University of Maryland<br />

at Baltimore, Baltimore, MD<br />

Bex proteins are a family of “Brain expressed X-linked genes” that<br />

are closely linked on the X-chromosome. Bex1 and 2 have been<br />

characterized as interacting partners of OMP. Here we report the<br />

development and characterization of an antibody to mouse Bex1 protein<br />

that also cross-reacts with Bex2 (but not Bex3) and its use to determine<br />

the first comprehensive distribution of Bex proteins in the murine brain.<br />

Immuno-blots and immunocytochemical analyses of cells transfected<br />

with either Bex1 or Bex2 have shown that the antiserum reacts with<br />

both Bex1 and Bex2. Antibodies preabsorbed with recombinant Bex2<br />

still recognize Bex1 while blocking with Bex1 totally eliminates all<br />

immunoreactivity to Bex1 and Bex2. Bex protein immunoreactivity (ir)<br />

was primarily localized to neuronal cells within select regions of the<br />

brain, including the olfactory bulb, epithelium, peri/paraventricular<br />

nuclei, suprachiasmatic nucleus, arcuate nucleus, median eminence,<br />

lateral hypothalamic area, thalamus, hippocampus, and cerebellum. Bex<br />

protein-ir was broadly present throughout the rostral-caudal aspects of<br />

the hypothalamic region. Further studies, using double-label<br />

immunocytochemistry, indicate that Bex-ir is colocalized with OMP in<br />

mature ORNs and in the OMP-positive subpopulation of hypothalamic<br />

neurons. This is the first anatomical demonstration of the<br />

comprehensive mapping of Bex proteins in the mouse brain and their<br />

colocalization with OMP in ORNs and hypothalamic neurons.<br />

Supported by NIH grants DC003112 and DC00054.<br />

102<br />

386 Poster [ ] Cell Biology of the Olfactory Epithelium<br />

REDUCED OLFACTORY EPITHELIUM MITOTIC RATE IN<br />

STREPTOZOTOCIN-INDUCED DIABETIC RATS<br />

Dennis J.C. 1, Swyers S. 2, Wright J.C. 3, Coleman E.S. 4, Judd R.L. 2, Hoe<br />

L. 2, Morrison E.E. 4 1Anatomy, Physiology and Pharmacology, Auburn<br />

University, Auburn, AL; 2Anatomy, Physiology, Pharacology, Auburn<br />

University, Auburn, AL; 3Pathobiology, Auburn University, Auburn,<br />

AL; 4Anatomy, Physiology, Pharmacology, Auburn University, Auburn,<br />

AL<br />

Individuals who suffer from diabetes frequently develop anosmia <strong>for</strong><br />

unknown reasons. The site(s) of disfunction in the olfactory pathways is<br />

(are) not known. We used immunohistochemistry to compare the<br />

numbers of mitotically active cells in the main olfactory epithelium<br />

(OE) of normal (n=4) male Wistar rats and in males (n=4) that suffered<br />

from streptozotocin (STZ)-induced diabetes. Insulin-dependent (type 1)<br />

diabetes was induced by intravenous injection of 50 mg/kg bw STZ.<br />

After 8 weeks, each animal was injected with bromodeoxyuridine<br />

(BrdU) (50ìg/gm bw). An hour later, the animals were deeply<br />

anesthetized with pentobarbital and perfused with buffered 4%<br />

para<strong>for</strong>maldehyde. After paraffin embedding, 7 ìm transverse sections<br />

were mounted on slides and hydrated. BrdU(+) cells were counted in 8<br />

nonserial sections from 3 regions dorsal to ventral per animal. The<br />

length of epithelium containing those cells was measured using Spot<br />

Advanced software. Counts were rendered as BrdU(+) cells/mm<br />

epithelium. Analysis by Student´s t Test using SAS 8.2 software<br />

indicated a significant difference in mitotic rates (p0.05).<br />

These data indicate that, after 8 weeks of STZ-induced type 1 diabetes<br />

in Wistar rats, mitotic rate in the main OE is lower in diabetics<br />

compared to normal controls.<br />

387 Poster [ ] Cell Biology of the Olfactory Epithelium<br />

QUALITATIVE AND QUANTITATIVE STUDY OF<br />

CYTOCHROME OXIDASE STAINING PATTERN IN<br />

OLFACTORY EPITHELIUM OF NEONATAL RAT<br />

Pataramekin P.P. 1, Meisami E. 1 1Molecular and Integrative Physiology,<br />

University of Illinois at Urbana-Champaign, Urbana, IL<br />

Newborn rats are capable of olfaction and their olfactory epithelium<br />

(OE) contains functionally mature olfactory receptor neurons (ORNs),<br />

although total ORN number is less than 10% of adult. Cytochrome<br />

oxidase (CO) staining density is positively correlated to levels of<br />

metabolic/neuronal activity. We stained newborn rat OE to characterize<br />

its staining pattern qualitatively and quantitatively. Staining<br />

densitometry was carried out using MCID software to gauge the<br />

luminescence of the sample over a range of 0.0 – 1.0 (higher numbers<br />

indicated lesser CO staining). A differential banding pattern of CO<br />

staining was observed corresponding to specific OE layers and ORN<br />

cellular sites, indicating the parts of the ORNs that were most actively<br />

involved in neuronal activity and transduction. Zones corresponding to<br />

ORN dendrite and knob exhibited high CO activity (0.34). Staining was<br />

heterogeneous within and along the length of dendrites, darker in the<br />

apical part, implying higher neuronal activity in this region than the<br />

deeper part of the dendrite, closer to the soma. A light staining band<br />

(0.40), between the two dark bands, corresponded to the overlapping<br />

presence of supporting cell cytoplasm and nuclei. Areas of OE close to<br />

the basal lamina also stained lightly (>0.44), implying lack of<br />

mitochondria and neuronal functional activity in the basal and immature<br />

neurons. The findings confirm that some ORNs are metabolically<br />

mature and functional in neonatal rats, capable of neuronal function.<br />

Support: University of Illinois Research Funds


388 Poster [ ] Cell Biology of the Olfactory Epithelium<br />

FUNCTIONAL CHARACTERIZATION OF CUB-SERINE<br />

PROTEASE IN THE SPINY LOBSTER'S OLFACTORY ORGAN<br />

Johns M. 1, Tai P. 1, Derby C.D. 1 1Biology, Georgia State University,<br />

Atlanta, GA<br />

Several serine proteases and protease inhibitors have been identified<br />

in the olfactory organ of decapod crustaceans (Hollins et al., 2003, J<br />

Comp Neurol 455:125-138; Stoss et al., 2002, Chem Senses 27:A45)<br />

including a CUB-serine protease (Csp: Levine et al., 2001, J Neurobiol<br />

49:277-302). The function of these proteases in the olfactory organ is<br />

unknown. To examine directly the functional activity of proteases in<br />

the olfactory organ of the Caribbean spiny lobster Panulirus argus, we<br />

used soluble and membrane tissue fractions from the lateral flagellum<br />

in a spectrophotometric enzyme activity assay with a variety of protease<br />

substrates and inhibitors to demonstrate trypsin-like serine protease<br />

activity. Csp accounts <strong>for</strong> at least 40% of the serine protease activity of<br />

the membrane fraction of the olfactory organ, based on<br />

immunoprecipitation of Csp. Serine protease activity follows a<br />

developmental pattern: it is lowest in the proximal zone, which lacks<br />

aesthetascs, and in the proliferation zone, where olfactory receptor<br />

neurons and associated cells are born, and highest in aesthetascs of the<br />

senescence zone, which has the oldest olfactory tissue. Csp activity is<br />

present at approximately the same level in each of the developmental<br />

zones. Currently we are determining if Csp´s functional activity<br />

changes due to damage of olfactory tissue. Supported by NIH grant<br />

DC00312.<br />

389 Poster [ ] Cell Biology of the Olfactory Epithelium<br />

EXOCRINE GLANDS CONTAINING SERINE PROTEASE ARE<br />

ASSOCIATED WITH OLFACTORY SENSILLA IN THE SPINY<br />

LOBSTER, PANULIRUS ARGUS<br />

Derby C.D. 1, Schmidt M. 1 1Biology, Georgia State University, Atlanta,<br />

GA<br />

In decapod crustaceans, the olfactory organ is constituted by<br />

specialized setae - aesthetascs - located on the lateral flagella of the 1st<br />

antennae. In an attempt to identify molecular components of the<br />

aesthetascs, diverse genes have been cloned from the lateral flagella of<br />

the spiny lobster, Panulirus argus, among them a gene encoding a<br />

CUB-serine protease (Csp) (Levine et al., J. Neurobiol. 49:277-302,<br />

2001). We found that an antibody against Csp specifically labels cells<br />

that are closely associated with but not a component of the aesthetascs.<br />

These cells are arranged in a rosette-like pattern of 5-7 cells around a<br />

duct with a terminal swelling revealed by labeling with phalloidin, an<br />

actin marker. The arrangement and phalloidin-reactivity are typical <strong>for</strong><br />

“rosette glands”, an exocrine epithelial gland of decapods (Talbot et al.,<br />

Zoomorphology, 110:329-338, 1991). SEM imaging revealed that in<br />

Panulirus argus the gland openings are a field of 50-70 dome-shaped<br />

cuticular structures (ca. 1 µm in diameter) with crescent-shaped pores<br />

proximal to each row of aesthetascs. Similar small pores have been<br />

found at the base of aesthetascs of other decapod crustaceans,<br />

suggesting that the association of exocrine rosette glands with<br />

aesthetascs is a common feature. From the presence of Csp among the<br />

excreted substances, we hypothesize that the rosette glands protect the<br />

aesthetascs from microbial fouling. Supported by NSF IBN-0077474<br />

and NIH DC00312<br />

103<br />

390 Poster [ ] Cell Biology of the Olfactory Epithelium<br />

FUNCTIONAL STUDIES OF A SERINE PROTEASE AND AN<br />

AMINE MONO-OXYGENASE SPECIFIC TO THE LOBSTER<br />

OLFACTORY ORGAN.<br />

Stepanyan R. 1, Day K.M. 1, Stepanyan T.D. 2, Mcclintock T.S. 1<br />

1Physiology, University of Kentucky, Lexington, KY; 2Molecular and<br />

Biomedical Pharmacology, University of Kentucky, Lexington, KY<br />

Olfactory specific proteins probably participate in critical functions<br />

of the olfactory system. We previously identified olfactory enriched<br />

transcript-03 (OET-03), which is expressed only by collar cells<br />

associated with the inner dendrites of the olfactory sensory neurons of<br />

Homarus americanus (Hollins et al., 2003. J. Comp. Neurol. 455:125).<br />

We now report the complete cDNA sequence of OET-03, a serine<br />

protease. Confirmation that OET-03 is a protease was obtained by<br />

assaying protease substrates and inhibitors. Removal of the presumed<br />

catalytic domain eliminated activity. Preliminary evidence indicates<br />

that OET-03 is functionally related to the chymotrypsin family. In<br />

addition, full-length OET-03 specifically reduces proliferation of<br />

transfected HEK293 cells. We also report the complete sequence of<br />

OET-02, expressed only by outer auxiliary cells ensheathing the<br />

proximal regions of the inner dendrites of the sensory neurons. OET-02<br />

has a unique double mono-oxygenase structure in which the two<br />

domains characteristic of dopamine beta-hydroxylases are repeated.<br />

Preliminary assays of olfactory tissue have not identified a biogenic<br />

amine product of this enzyme, so tests with recombinant OET-02 are<br />

planned. An antiserum raised against OET-02 confirms the large size of<br />

this protein (155 kDa) and its restricted expression to the outer auxiliary<br />

cells.<br />

Supported by R01 DC02366.<br />

391 Poster [ ] Olfactory Bulb: Neurophysiology<br />

QUANTIFICATION OF TAURINE-SYNTHESIZING ENZYME<br />

MRNA IN OLFACTORY STRUCTURES WITH REAL-TIME<br />

RT-PCR<br />

Kratskin I. 1, Hao Y. 1 1Smell and Taste Center, University of<br />

Pennsylvania, Philadelphia, PA<br />

The olfactory epithelium (OE) and olfactory bulb (OB) are rich in the<br />

amino acid taurine, which is synthesized in liver and nervous tissue via<br />

the enzyme cysteine sulfinate decarboxylase (CSD). Our previous<br />

studies have shown that (i) taurine, acting on GABA receptors, inhibits<br />

mitral/tufted cells and their synaptic responses to olfactory nerve input,<br />

(ii) immunoreactivities <strong>for</strong> taurine and CSD are located in olfactory<br />

receptor cells and most OB neurons, and (iii) CSD mRNA is expressed<br />

in the OE and OB. In the present study, gene expression of CSD in the<br />

rat OE and OB was assessed quantitatively using the real time reverse<br />

transcription-polymerase chain reaction (RT-PCR) in the LightCycler.<br />

With this method, a PCR product is detected after each amplification<br />

cycle by measuring fluorescence emission of a reporter dye, whose<br />

release is proportional to the amount of the product. A calibration curve<br />

(a regression line with a correlation coefficient of 0.99) showing the<br />

relationship between a starting copy number of CSD mRNA and a<br />

threshold cycle number within a range of 200 to 20000,000 mRNA<br />

copies was obtained first. Values of the threshold cycle number <strong>for</strong><br />

tissue samples were then detected and a copy number of CSD mRNA<br />

was calculated using a regression equation. The level of CSD mRNA in<br />

both OE and OB was about 1000,000 copies/mg tissue. A higher<br />

amount of about 100,000,000 CSD mRNA copies/mg tissue was<br />

detected in the liver. The results confirm that CSD mRNA is expressed<br />

in the OE and OB and provide its quantitative evaluation. (Supported<br />

by NIDCD grant DC04083).


392 Poster [ ] Olfactory Bulb: Neurophysiology<br />

KV1.3-NULL MUTATION ALTERS SCAFFOLDING<br />

PROTEINS, OLFACTORY BULB BIOPHYSICS, AND<br />

GLOMERULI SIZE/ABUNDANCE.<br />

Perkins R.M. 1, Tucker K. 2, Meredith M. 2, Fadool D. 2 1Dept. of<br />

Biological Science, Florida State University, Tallahassee, FL; 2Prog. in<br />

Neurosci. & Mol. Biophys., Florida State University, Tallahassee, FL<br />

Previously we demonstrated that Kv1.3-null (KO) mice had smaller<br />

sized and more abundant glomeruli independent of coronal position in<br />

the olfactory bulb (OB). To test whether projections to the OB were<br />

altered, we generated Kv1.3-/- mice with a targeted P2-IRES-taulacZ<br />

mutation (KvP2). Coronal cryosections (12 µm) and whole mounts of<br />

OB and olfactory epithelium of KvP2 and P2 mice (n=3-5) were<br />

photographed using a Zeiss Axiocam digital camera with Axiovision<br />

software. The location of P2 receptors along the olfactory epithelium<br />

and known P2 glomeruli in the OB showed no noticeable differences<br />

between P2 and KvP2 mice. In spite of a demonstrated lack of change<br />

in axonal connectivity, action potentials recorded from mitral cells of<br />

KO mice exhibited increased hyperpolarization, 10-90% rise time, and<br />

width at ½ maximum amplitude. Current injection elicited higher<br />

spiking frequency with less variance in interpulse interval. OB neurons<br />

are no longer modulated by activation of receptor tyrosine kinases<br />

(insulin or TrkB receptor) typically associated with the Kv1.3 channel.<br />

SDS-PAGE indicates the upregulation of Ca2+ channels, two tyr<br />

kinases, and six adapter proteins in the OB of KO mice. The electrical<br />

properties in mitral cells from Kv1.3 KO mice could produce rapid<br />

synchronous firing to strengthen connections to the more numerous<br />

glomeruli. This interpretation would be consistent with other<br />

experiments in our laboratory that indicate an increased olfactory ability<br />

in the mice. Support: NIH DC03387 (NIDCD) & Florida Foundation.<br />

393 Poster [ ] Olfactory Bulb: Neurophysiology<br />

MULTIPLE ROLES OF TRKB RECEPTOR IN MODULATING<br />

KV1.3 ION CHANNEL IN THE OLFACTORY BULB.<br />

Colley B.S. 1, Visegrady A. 1, Fadool D. 1 1Prog. in Neurosci. & Mol.<br />

Biophysics, Florida State University, Tallahassee, FL<br />

Previously we have shown that olfactory bulb neuron (OBN) current<br />

properties are modulated by activation of the neurotrophin receptor<br />

TrkB, by brain-derived neurotrophic factor (BDNF), to increase<br />

tyrosine phosphorylation of the potassium channel Kv1.3. Here we<br />

made Y to F mutations in the Kv1.3 channel to test whether removal of<br />

the Y phosphorylation motif disrupted BDNF-induced current<br />

suppression in Kv1.3 + TrkB transfected HEK293 cells. Tyrosines in<br />

the N and C termini (111-113, 137, 449) were found to be the molecular<br />

targets <strong>for</strong> current modulation via phosphorylation as demonstrated by<br />

patch-clamp electrophysiology and phosphorylation assays using<br />

various mutant channel constructs compared with the wildtype channel.<br />

Unexpectedly, we found a 2.1 fold increase in channel protein<br />

expression (n=4) and a 2 fold increase in peak current amplitude in<br />

HEK293 cells co-transfected with channel + TrkB versus channel<br />

alone. This phosphorylation-independent upregulation of Kv1.3 was<br />

not observed using a GFP-tagged Kv1.3 that likely disrupts an Nterminal<br />

ER retention motif. Kv1.3 can be immunoprecipitated with<br />

TrkB in the presence of the adaptor protein nShc and the described<br />

upregulation is disrupted by substituting trkB K538N or Y490F, which<br />

are trkB mutants that lack kinase activity or association with nShc<br />

adaptor respectively. These results suggest that TrkB receptor may<br />

alter the excitability of OBNs by both phosphorylation-dependent and -<br />

independent mechanisms involving Kv1.3 ion channel.<br />

Supported by NIH R01 DC03387 (NIDCD).<br />

104<br />

394 Poster [ ] Olfactory Bulb: Neurophysiology<br />

GABAERGIC PERIGLOMERULAR CELLS<br />

PRESYNAPTICALLY INHIBIT ON INPUT TO THEMSELVES<br />

Shao Z. 1, Szabo G. 2, Puche A.C. 1, Shipley M.T. 1 1University of<br />

Maryland at Baltimore, Baltimore, MD; 2Dept. of Gene Tech. and Dev.<br />

Neurobiol., Institute of Exp. Med., Budapest, ., Hungary<br />

Olfactory nerve axons terminate in olfactory bulb glomeruli and <strong>for</strong>m<br />

excitatory synapses onto the dendrites of mitral/tufted (M/T) and<br />

periglomerular (PG) cells. ON terminals express dopamine D2 and<br />

GABAB receptors and there is evidence that DA and GABA<br />

presynaptically inhibit the release of glutamate from ON synapses. It is<br />

thought that ON excitation of GABAergic PG cells releases GABA<br />

back onto ON terminals, but there is no direct evidence <strong>for</strong> this. To<br />

investigate this question we used whole cell patch clamp recordings of<br />

identified GABAergic PG cells in a line of mice expressing green<br />

fluorescent protein under the control of the glutamic acid decarboxylase<br />

65kDa gene (GAD65-GFP). GAD65-GFP+ GABAergic PG cells are<br />

excited by ON stimulation; 30%-40% receive monosynaptic ON input,<br />

the remainder have variable latency, polysynaptic responses to ON<br />

stimulation. For those that receive monosynaptic input, paired-pulse<br />

ON stimulation caused a reduction of EPSC amplitude in response to<br />

the second pulse. This paired-pulse inhibition was abolished by addition<br />

of selective GABAB receptor antagonists. This indicates that ON<br />

terminals targeting GABAergic PG cells are subject to GABAergic<br />

presynaptic inhibition. When step-depolarization of the cell was used to<br />

evoke GABA release prior to stimulation of ON, the magnitude of the<br />

ON-evoked EPSC was reduced by ~30%. The reduction was prevented<br />

by application of selective GABAB receptor antagonists. This<br />

demonstrates that a GABAergic PG cell can presynaptically inhibit the<br />

ON terminals targeting that same PG cell. GABAergic PG cells,<br />

there<strong>for</strong>e, may provide feedback, presynaptic inhibition that limits<br />

subsequent glutamate release from ON terminals. Supported by NIH<br />

DC02173, DC00347 & NS36940.


395 Poster [ ] Olfactory Bulb: Neurophysiology<br />

INTRAGLOMERULAR SYNCHRONOUS CALCIUM<br />

OSCILLATIONS OF PERIGLOMERULAR CELLS IN THE<br />

MOUSE OLFACTORY BULB<br />

Jia F. 1, Shepherd G.M. 1, Chen W.R. 1 1Neurobiology, Yale University,<br />

New Haven, CT<br />

Synchronous neuronal oscillations have been widely considered to be<br />

an important mechanism <strong>for</strong> coding and processing odor in<strong>for</strong>mation.<br />

Oscillations of different frequency ranges have been reported in the<br />

olfactory bulbs of a wide range of vertebrate species. Olfactory sensory<br />

neurons expressing the same odorant receptor project to a few<br />

glomeruli in the mammalian olfactory bulb. Thus, glomeruli are key<br />

<strong>for</strong> the neural basis of olfactory coding. In this study we have used<br />

optical imaging methods to map neuronal synchrony among<br />

periglomerular (PG) cell populations that either share a common<br />

glomerulus or belong to different glomeruli. An ester <strong>for</strong>m of calciumsensitive<br />

dye was used to load hundreds of neurons in the olfactory bulb<br />

slices. Thirty to one hundred fifty seconds of calcium imaging was<br />

per<strong>for</strong>med to analyze the dynamic structure of PG-cell spontaneous<br />

activities by a cooled CCD camera. We observed novel slow (


398 Poster [ ] Olfactory Bulb: Neurophysiology<br />

OLFACTORY NERVE-EVOKED METABOTROPIC<br />

GLUTAMATE RECEPTOR (MGLUR)-MEDIATED<br />

RESPONSES IN RAT OLFACTORY BULB MITRAL CELLS<br />

Zhu M. 1, Heinbockel T. 2, Ennis M. 1 1Anatomy and Neurobiology,<br />

University of Tennessee, Memphis, TN; 2Anatomy and Neurobiology,<br />

University of Maryland at Baltimore, Baltimore, MD<br />

The group I mGluR, mGluR1, is highly expressed on mitral cell<br />

(MC) apical dendrites, and thus may be activated by glutamate from the<br />

olfactory nerve (ON) and/or mitral/tufted cells. We have shown that<br />

mGluR1 agonists potently excite MCs. Here we investigated if ON<br />

stimulation engages mGluR-mediated responses in MCs in rat olfactory<br />

bulb slices. In voltage clamp recordings, EPSCs evoked by single ON<br />

shocks were unaffected by the potent, non-selective mGluR antagonist<br />

LY341495 (50-100 uM). Single ON shocks in the presence of<br />

ionotropic glutamate and GABAa receptor antagonists (CNQX, APV,<br />

Gabazine) did not evoke EPSCs, except at high intensities (>600uA). In<br />

the same cell (in CNQX, APV, Gabazine), brief high frequency ON<br />

trains (6 pulses at 10-200Hz) induced long duration EPSCs that were<br />

blocked by TTX or by low calcium ACSF. Application of glutamate<br />

uptake inhibitor (TBOA 100uM) and a glutamate transporter inhibitor<br />

(THA 300uM) increased the amplitude and duration of responses to<br />

single or trains of ON stimuli. These agents also unmasked “latent”<br />

responses to single ON stimuli in some cells. Additional application<br />

LY341495 (50-100uM) reduced or completely blocked ON-evoked<br />

responses observed in the presence CNQX, Gabazine and APV. These<br />

results demonstrate that ON stimulation evokes frequency dependent<br />

excitatory responses in MCs that are mediated, in part, by activation of<br />

mGluRs. Such responses are maximally engaged during high frequency<br />

ON input, as occurs during odor stimulation. Support: PHS grants<br />

DC03195, DC00347<br />

399 Poster [ ] Olfactory Bulb: Neurophysiology<br />

METABOTROPIC GLUTAMATE RECEPTORS (MGLURS)<br />

ENHANCE SYNAPTIC INTERACTIONS AMONG<br />

JUXTAGLOMERULAR (JG) NEURONS IN OLFACTORY<br />

BULB GLOMERULI<br />

Hayar A. 1, Heinbockel T. 2, Shipley M.T. 2, Ennis M. 1 1Dept. of Anatomy<br />

and Neurobiology, University of Tennessee, Memphis, TN; 2Dept. of<br />

Anatomy and Neurobiology, University of Maryland, Baltimore, MD<br />

mGluRs are present on JG cells and may regulate dendrodendritic<br />

synapses. We investigated the actions of mGluR agonists in identified<br />

JG cells in slices. Group I (DHPG), and group II (CCGI) mGluR<br />

agonists, but not group III mGluR agonists (L-AP4) increased the<br />

number of spikes/burst in external tufted (ET) cells in the presence of<br />

synaptic blockers. In the presence of TTX and synaptic blockers,<br />

DHPG and CCGI directly excite and induce an inward current in ET<br />

cells. They also increased the frequency of spontaneous and miniature<br />

IPSCs in ET cells. This effect, abolished by the Ca2+ channel blocker,<br />

cadmium, may be due to increased ET and/or mitral cell dendritic<br />

excitability and/or a direct, mGluR-mediated excitatory effect on the<br />

dendrites of GABAergic periglomerular (PG) cells. DHPG and CCGI<br />

also increased the frequency of miniature EPSCs in PG and short axon<br />

(SA) cells but not in ET cells, suggesting that activation of mGluRs on<br />

ET cell dendrites, increases glutamate release onto PG and SA cells. In<br />

the presence of fast synaptic blockers and the Na + channel blocker QX-<br />

314 in the pipette, DHPG induced in ET cells rhythmic events<br />

associated with membrane current oscillations (~2 Hz) that are possibly<br />

due to gap junctions among ET cells. mGluR agonists also decreased<br />

olfactory nerve-evoked EPSCs in ET and mitral cells via a GABA - B<br />

mediated presynaptic mechanism. Thus, mGluR agonists directly<br />

activate rhythmically bursting ET cells and enhance glomerular<br />

dendrodendritic interactions. Support: PHS Grants: DC03195,<br />

DC05676.<br />

106<br />

400 Poster [ ] Olfactory Bulb: Neurophysiology<br />

IN VIVO MOUSE PREPARATION FOR OLFACTORY BULB<br />

ELECTROPHYSIOLOGY<br />

Mast T. 1, Griff E. 2 1Biological <strong>Sciences</strong>, University of Cincinnati*,<br />

Cincinnati, OH; 2Biological <strong>Sciences</strong>, University of Cincinnati,<br />

Cincinnati, OH<br />

The objective of this study was to develop an in vivo mouse<br />

preparation <strong>for</strong> recording from the main olfactory bulb (MOB). An aim<br />

was to establish a protocol using an injectable anesthetic such as chloral<br />

hydrate. The anesthetic plane was evaluated using the EEG and the<br />

hindpaw withdrawal response. Single-unit recordings measured the<br />

spontaneous activity of neurons in the MOB. The amount of chloral<br />

hydrate required to maintain a surgical plane of anesthesia was linearly<br />

correlated with a decrease in MOB spontaneous activity (p=0.2; n=13),<br />

and could be lethal. There<strong>for</strong>e, analgesic supplements to chloral<br />

hydrate anesthesia were investigated. Supplementation with<br />

buprenorphine, a synthetic µ-opioid receptor agonist, was problematic.<br />

Buprenorphine at doses of 0.02, 0.05, and 0.2 mg/kg reduced<br />

spontaneous activity of bulbar neurons by 19, 45, and 54%,<br />

respectively. In contrast, ketoprofen, a non-steroidal anti-inflammatory<br />

with reported analgesic activity, did not inhibit spontaneous activity at<br />

doses of 100 or 200 mg/kg. Importantly, mice given 100 mg/kg<br />

ketoprofen at the beginning of an experiment, be<strong>for</strong>e surgery, required<br />

significantly less chloral hydrate than control animals (p=0.05; n=18).<br />

Animals in both groups were maintained at a surgical plane of<br />

anesthesia. Lastly, administration of ketoprofen during an experiment<br />

changed the EEG indicative of a deeper plane of anesthesia. Thus,<br />

chloral hydrate supplemented with ketoprofen in the mouse provides an<br />

in vivo preparation <strong>for</strong> stable, long-term recording of neurons in the<br />

MOB. This work was supported in part by NIH grant 1R15DC04548 to<br />

ERG.<br />

401 Poster [ ] Olfactory Bulb: Neurophysiology<br />

SPONTANEOUS ACTIVITY OF MAIN OLFACTORY BULB<br />

NEURONS IN THE RAT<br />

Griff E.R. 1, Stakic J. 2, Suchanek J. 2 1Biological <strong>Sciences</strong>, University of<br />

Cincinnati*, Cincinnti, OH; 2Biological <strong>Sciences</strong>, University of<br />

Cincinnati, Cincinnati, OH<br />

Neurons in the main olfactory bulb (MOB) are spontaneously active,<br />

with rates above 30 Hz reported. In this study, the sources of this<br />

activity were investigated. We specifically tested the hypothesis that<br />

the spontaneous activity of bulbar neurons is at least partly due to<br />

spontaneous activity in olfactory receptor neurons (ORNs).<br />

This study was conducted using anesthetized rats breathing clean air.<br />

The conduction of ORN action potentials was blocked by cooling the<br />

olfactory nerve rostral to the cribi<strong>for</strong>m plate. To assess the efficacy of<br />

the cold block, a bipolar stimulation electrode was positioned rostral to<br />

the cold probe, and field potentials were recorded from the MOB while<br />

cooling the nerve. Cooling completely and reversibly abolished the<br />

field potentials. The spontaneous activity of single units was recorded<br />

from the same region of the MOB, and the recording sites subsequently<br />

marked with dye. Mitral and tufted cells were further identified by<br />

antidromic activation from the lateral olfactory tract. To date, <strong>for</strong> every<br />

cell recorded that exhibited spontaneous activity (from glomerular,<br />

external plexi<strong>for</strong>m, and mitral cell layers of the MOB), cold decreased<br />

this activity. This study suggests that ORNs are spontaneously active in<br />

the rat. Thus, in the absence of odor stimulation, ORN spontaneous<br />

activity, directly and/or via bulbar circuits, at least in part sets and<br />

maintains neuronal activity in the MOB.<br />

This work was supported in part by NIH grant 1R15DC04548 to<br />

ERG.


402 Poster [ ] Olfactory Bulb: Neurophysiology<br />

MULTIUNIT AND FIELD POTENTIAL RECORDINGS IN RAT<br />

OLFACTORY BULB<br />

Sherrill L. 1, Green E. 1, Scott J.W. 1 1Cell Biology, Emory University,<br />

Atlanta, GA<br />

To test the relationship between olfactory bulb slow waves and single<br />

cell responses, we conducted recordings with silicon based multiple<br />

electrodes. Field potential recordings are used to calculate one and twodimensional<br />

current source density in rats anesthetized with urethane<br />

(1.5 gm/K) or chloral hydrate (0.4/K). Antidromic stimulation of the<br />

lateral olfactory tract is used to localize the mitral cell layer and in some<br />

cases to identify output cells. Under both anesthetics a number of<br />

odorants can evoke spike responses and waves in the 30-60 Hz range at<br />

concentrations in the range of 10-3 times saturated vapor in our<br />

preliminary results. The field responses evoked by odorants are more<br />

obvious in the voltage records than in the current density calculations.<br />

This agrees with our findings that the voltage oscillations are not well<br />

localized in the olfactory bulb. We failed to find spikes in either the<br />

mitral or granule layers that are strongly synchronized with slow waves.<br />

On the other hand it has been common to see strong inhibition or<br />

excitation of spikes in the mitral cell layer during bursts of oscillations.<br />

While some of these cells may be too strongly inhibited to show<br />

oscillations, preliminary attempts to uncover the underlying oscillation<br />

by reducing odorant concentration have been unsuccessful. While the<br />

prominence of these oscillatory phenomenia in the bulb suggests that<br />

they represent a strong synchronization of inibitory interneurons, our<br />

results suggest that it will be difficult to show simple relationships with<br />

spike activity. Supported by NIH grant DC00113. Multichannel silicon<br />

probes were provided by the Univ.Mich. Center <strong>for</strong> Neural<br />

Communication Technology sponsored by NIH NIBIB grant P41-<br />

RR09754.<br />

403 Poster [ ] Olfactory Bulb: Neurophysiology<br />

NITRIC OXIDE MODULATES ANTENNAL LOBE NEURON<br />

ACTIVITY IN THE MOTH, MANDUCA SEXTA, THROUGH<br />

SOLUBLE GUANYLYL CYCLASE-DEPENDENT<br />

MECHANISMS<br />

Wilson C. 1, Christensen T. 1, Nighorn A. 1 1Neurobiology, University of<br />

Arizona, Tucson, AZ<br />

Gaseous messengers like nitric oxide (NO) have been implicated in a<br />

number of physiological processes including the modulation of<br />

olfactory representations in the brain. NO and its signaling components<br />

have been studied in the olfactory systems of several species, but the<br />

function of NO in olfactory processing remains elusive. In the moth,<br />

Manduca sexta, NO synthase (NOS) was found in olfactory receptor<br />

neurons (ORNs), while a well-characterized target of NO, soluble<br />

guanylyl cyclase (sGC) was found in the post-synaptic targets of ORNs<br />

in antennal lobe (AL). This expression pattern, along with preliminary<br />

imaging data showing increases in fluorescence of an NO-sensitive dye<br />

in odor-activated glomeruli, led to a hypothesis that NO is released<br />

focally when odor is present, and then modulates signaling in the<br />

activated glomeruli by acting on the sGC-containing neurons in the AL.<br />

This hypothesis was tested with multi-channel and intracellular<br />

recording methods coupled with pharmacological manipulation of NO<br />

levels and the activity of sGC in the AL. Blocking NO production with<br />

L-NAME resulted in an overall decrease of spontaneous activity in AL<br />

neurons, and this effect was mimicked reversibly by the sGC inhibitor<br />

ODQ. Odor-evoked responses were also diminished when NOdependent<br />

signaling was blocked. These results suggest that NO is an<br />

important modulator of odor responsiveness, and that sGC-dependent<br />

signaling is at least partially responsible <strong>for</strong> the observed changes in<br />

glomerular activity levels. Supported by NIH-NIDCD DC04292 and<br />

DC06368.<br />

107<br />

404 Poster [ ] Multimodal Integration<br />

BEHAVIORAL RESPONSES OF THE CRAYFISH<br />

PROCAMBARUS CLARKII TO SINGLE COMPOUNDS<br />

Corotto F.S. 1, Johnston M.E. 1, Rogers J.L. 1, Williams J.M. 1<br />

1Department of Biology, North Georgia College & State University,<br />

Dahlonega, GA<br />

We investigated chemosensory behavior of the crayfish Procambarus<br />

clarkii in response to glutamate and glucose (not stimulatory to leg<br />

chemoreceptors in physiological tests; J. Chem. Ecol., 28:1117-1130),<br />

ammonium, glycine, maltose, and trehalose (physiological stimuli of<br />

varying efficacies), and a blank. Each animal was acclimated to a 1.5 L<br />

testing chamber <strong>for</strong> >2 h and was stationary in the chamber´s shelter<br />

when tested. Individual compounds were injected in 10 mL aliquots<br />

into a 149 mL/min carrier flow and were diluted to 200 µM-2 mM<br />

within the testing chamber according to conductivity measurements.<br />

Each animal was tested once. Responses were filmed, and the time<br />

spent probing with pereopods, the time spent exploring, and the number<br />

of dactyl clasps were quantified by observers blind to the stimulus.<br />

ANOVA followed by Fisher´s PLSD tests indicated the following<br />

elicited responses that differed significantly from that evoked by the<br />

blank: glucose <strong>for</strong> the time spent probing; glucose, maltose, and glycine<br />

<strong>for</strong> exploration time; glucose, maltose, and trehalose <strong>for</strong> the number of<br />

dactyl clasps. Ammonium´s ineffectiveness parallels its weakness as a<br />

physiological stimulus. Trehalose was the most potent physiological<br />

stimulus tested, but it was no more effective in evoking behavioral<br />

responses than glucose or maltose. Glucose was a potent behavioral<br />

stimulus, but earlier physiological tests failed to show responses to<br />

glucose from leg chemoreceptor cells. Either receptor cells <strong>for</strong> glucose<br />

are elsewhere or they are on the legs in small numbers and carry a<br />

disproportionate influence on behavior.<br />

405 Poster [ ] Multimodal Integration<br />

DUAL INTRACELLULAR RECORDINGS FROM SINGLE<br />

PARASOL CELLS REVEAL IMPULSE BURST INITIATION<br />

SITE AND DENDRITIC TRAJECTORIES<br />

Mellon D. 1 1Biology, University of Virginia, Charlottesville, VA<br />

Parasol cells are multimodal sensory interneurons in the crustacean<br />

<strong>for</strong>ebrain. They exhibit complex and interesting <strong>for</strong>ms of electrical<br />

activity, investigations of which may reveal novel concepts of dendritic<br />

integrative mechanisms. We are using dual intracellular recordings<br />

from individual parasol cells to examine impulse and impulse burst<br />

initiation sites and propagation within the dendritic tree. Previous<br />

findings (Mellon, DeF., J. Neurophysiol. 90:2465 [2003]) indicated that<br />

impulses are initiated at a low threshold region on the dendritic trunk<br />

adjacent to the basal branch points. Dual simultaneous recordings from<br />

different basal branches of the same neuron support this interpretation<br />

and, furthermore, suggest that impulse bursts arise at a trunk region<br />

distant from (more ventral to) the intitation zone of single spikes.<br />

Orthodromic stimulation via the olfactory-globular tract or antidromic<br />

stimulation via the parasol cell axon terminals generates slightly<br />

different impulse latencies in simultaneous recordings from basal<br />

branches on opposite sides of the dendritic tree. Impulse bursts, evoked<br />

either by photic stimulation of the compound eyes or by posthyperpolarization-activated<br />

I h , invade both recording sites and exhibit<br />

latency differences identical with those of individual spikes following<br />

orthodromic or antidromic stimulation. These findings support<br />

conclusions regarding a trunk site <strong>for</strong> impulse initiation and provide<br />

evidence <strong>for</strong> bursts also being initiated on the dendritic trunk.<br />

Supported by NSF.


406 Poster [ ] Multimodal Integration<br />

TUNING OF MECHANORECEPTORS IN THE LOBSTER<br />

ANTENNULE<br />

Miller-Sims V. 1, Atema J. 1 1Biology, Boston University, Woods Hole,<br />

MA<br />

The American lobster, Homarus americanus, is capable of tracking<br />

an odor to its source in a turbulent plume. In such a plume an animal<br />

encounters hydrodynamic and chemical signals. Concurrent reception<br />

of odor and flow could be an important cue <strong>for</strong> odor plume tracking.<br />

The lateral antennule contains many chemoreceptor neurons and is<br />

important <strong>for</strong> successful tracking. We now want to determine the<br />

bandwidth in which the lobster uses lateral antennule mechanoreceptors<br />

to monitor the hydrodynamic properties of the plume. Under oscillatory<br />

flow conditions the antennule has a resonance of frequency of 5-12 Hz;<br />

antennular mechanoreceptors could also be expected to respond<br />

optimally in this range. In this experiment frequency synchronization is<br />

used to determine the types of hydrodynamic stimuli to which the<br />

antennule is capable of responding. The distal end of the antennule was<br />

fixed in a speaker driven oscillatory flow chamber to stimulate whole<br />

antennule movement. We recorded extracellular responses from<br />

antennular axons to oscillatory antennule movement over a frequency<br />

range of 1-128 Hz in an amplitude range of 16 to125 µm. Responses<br />

were recorded from 30 cells and synchronization coefficients were<br />

computed <strong>for</strong> each stimulus presentation. As a population these cells<br />

were capable of synchronizing to the stimulus at frequencies between 4<br />

and 64 Hz. Cells with low to no background spiking synchronized best<br />

at lower frequencies of 4-16 Hz while cells with high background<br />

spiking synchronized best at higher frequencies between 16 and 64 Hz.<br />

This frequency range may have significance in monitoring turbulent<br />

plume dynamics. Funding Source: NSF IBN-0091358<br />

407 Poster [ ] Multimodal Integration<br />

OCTOPAMINE-IMMUNOREACTIVE NEURONS IN THE<br />

OLFACTORY AND GUSTATORY CENTERS IN THE BRAIN<br />

OF MANDUCA SEXTA.<br />

Dacks A. 1, Christensen T. 1, Hildebrand J.G. 2 1Neurobiology, University<br />

of Arizona, Tucson, AZ; 2Univeristy of Arizona*, Tucson, AZ<br />

Octopamine (OA) is a neuroactive monoamine that is found in the<br />

nervous systems of many invertebrates and plays a key role in different<br />

aspects of behavior, including associative learning. In this study, we<br />

used a monoclonal antibody against OA (gift of H.-J. Agricola) to<br />

examine the distribution of OA-immunoreactive (OA-ir) cells in the<br />

olfactory and gustatory neuropil centers that mediate olfactory<br />

conditioning in the hawkmoth Manduca sexta. OA-ir processes were<br />

observed in many regions of the brain with the distinct exception of the<br />

upper division of the central body. We found 11 ventral unpaired<br />

median (VUM) cell bodies in the subesophageal ganglion that express<br />

OA-ir. Two particularly elaborate cells project bilaterally: one<br />

interconnects the antennal lobes and the calyces (input region) of the<br />

mushroom bodies, and a second cell innervates the gamma-lobe of the<br />

mushroom bodies. The branching patterns of Manduca VUM neurons<br />

are there<strong>for</strong>e similar, but not identical to those in other insects,<br />

including the uniquely identifiable VUMmx1 neuron in the honeybee.<br />

Supported by grants from NIH/NIDCD and NSERC Canada.<br />

108<br />

408 Poster [ ] Multimodal Integration<br />

TEMPORAL INTERACTION OF OLFACTORY AND<br />

TRIGEMINAL PROCESSING<br />

Heilmann S.K. 1, Hummel T. 1 1Otorhinolaryngology, University of<br />

Dresden Medical School, Dresden, Germany<br />

Aim:<br />

In everyday life, olfactory and trigeminal sensations hardly ever<br />

occur seperately. Contrary, most odors in our environment will elicit<br />

excitation of both olfactory receptor cells and trigeminal neurons.<br />

However, the temporal resolution of this interaction has not been<br />

studied in detail yet, mostly due to technical difficulties. Aim of this<br />

study was to analyse the interaction of both systems using intensity<br />

estimates following olfactory and trigeminal stimuli with different<br />

interstimulus intervals.<br />

Methods:<br />

Participants were 11 women and 12 men (age 21-29 years).<br />

Trigeminal (50 % v/v CO2) and olfactory (5 ppm H2S) stimuli were<br />

presented at different interstimulus intervals (0, 700, 1200, 2200, 4200<br />

and 8200 ms). One group of subjects received first the olfactory<br />

stimulus, followed by a trigeminal stimulation (OT). A second group<br />

received the stimuli in reversed order (TO). Intensity estimates were<br />

recorded using a visual analogue scale.<br />

Results:<br />

OT: Olfactory stimulation resulted in a significant enhancement of<br />

perceived intensity of the following trigeminal stimulus <strong>for</strong> up to 1000<br />

ms. When the stimulation sequence was reversed (TO), perceived<br />

intensity of the olfactory stimulus was significantly reduced <strong>for</strong> up to<br />

2000 ms.<br />

Summary:<br />

With regard to the present experimental conditions, the olfactory and<br />

trigeminal systems interact within a time frame of 1-2 s. It appears that<br />

trigeminal activation affects the processing of other chemosensory<br />

stimuli <strong>for</strong> a longer period than olfactory excitation.<br />

409 Poster [ ] Multimodal Integration<br />

PROP TASTER STATUS VERSUS TEXTURE AND FLAVOR<br />

SENSATIONS FOR LOW-FAT SEMI-SOLID FOODS<br />

De Wijk R.A. 1, Weenen H. 2 1Wageningen Center <strong>for</strong> Food<br />

<strong>Sciences</strong>/A&F, Wageningen, Netherlands; 2WCFS/TNO-V,<br />

Wageningen, Netherlands<br />

The hypothesis was tested that sensitivity to 6-n-propylthiouracil<br />

(PROP) was related to texture and flavor perception, and preference <strong>for</strong><br />

16 commercial vanilla custard desserts (mostly starch-based, fat levels<br />

between 0.1% and 4%). A group of 175 naive consumers rated the<br />

intensity of a PROP stimulus presented on a filter paper using a gLMS.<br />

The group was divided into 25% non-tasters (NT, rating=86%). PROP status affected ratings <strong>for</strong> creaminess, fattiness,<br />

thickness, melting, heterogeneity, airiness, stickiness, and roughness,<br />

and preference significantly (p


410 Poster [ ] Multimodal Integration<br />

EFFECTS OF GINKGO BILOBA ON CHEMOSENSORY<br />

FUNCTION<br />

Mattes R.D. 1, Pawlik K. 2 1Foods and Nutrition, Purdue University,<br />

West Lafayette, IN; 2Foods and Nutrition, Purdue University, W.<br />

Lafayette, IN<br />

Diminutions and losses of chemosensory function are common.<br />

While diagnostic procedures have advanced markedly, there are few<br />

known methods to enhance chemosensory function. Ginko biloba<br />

reportedly enhances cerebral blood flow, neurogenesis in the olfactory<br />

epithelium, neurotransmission and cognition. Improved odor<br />

recognition has been noted in rats administered the extract. This trial<br />

examined the effects of acute and chronic extract administration on<br />

taste and smell threshold sensitivity and odor recognition in healthy<br />

humans. A randomized, double-blind, placebo-controlled trial was<br />

underetaken with 30 individuals meeting strict eligibility criteria<br />

including non-use of medications, nutraceuticals or supplements that<br />

can affect sensory function. Active treatment consisted of a daily<br />

0.87mg/kg dose (2.35% Ginkgo biloba extract standardized to 24%<br />

flavanoids and 6% terpenes (%w/w)) administed TID <strong>for</strong> 13 weeks.<br />

Forced-choice, staircase thresholds <strong>for</strong> sucrose, NaCl and butanol and<br />

odor identification were measured at weeks 0, 1, 5, 9 and 13 following<br />

a standardized meal. No significant treatment or treatment X time<br />

interactions were observed <strong>for</strong> the olfactory indices. Taste thresholds<br />

were significantly higher with active treatment over the trial. This trial<br />

revealed no beneficial effects of Ginkgo biloba on taste or smell in<br />

healthy adults. Whether benefits may be realized by individuals with<br />

diminished chemosensory function is not known.<br />

Supported by: Purdue-UAB Botanicals Center <strong>for</strong> Dietary<br />

Supplement Research #P50 AT-00477.<br />

411 Poster [ ] Multimodal Integration<br />

INFLUENCES OF ANTIHYPERTENSIVE AND<br />

ANTIHYPERLIPIDEMIC DRUGS ON THE SENSES OF TASTE<br />

AND SMELL: AN OVERVIEW<br />

Kerr K.L. 1, Philip S. 1, Reddy K. 1, Doty R.L. 1 1Smell and Taste Center,<br />

Department of Otorhinolaryngology, University of Pennsylvania<br />

Medical Center, Philadelphia, PA<br />

According to the Physician´s Desk Reference (PDR), 36% of modern<br />

antihypertensive and antihyperlipidemic drugs produce untoward<br />

alterations in chemosensory perception. Such disturbances can<br />

adversely affect the quality of life, produce noncompliance to<br />

medication schedules, and may result in decreased food intake, loss of<br />

appetite, weight decrement, and depression. This abstract lists the<br />

primary antihypertensive and antihyperlipidemic drugs that adversely<br />

alter chemosensory function. Among these, angiotensin-converting<br />

enzyme (ACE) inhibitors and antihyperlipidemic drugs are the worst<br />

offenders, as seven of the 10 (70%) ACE inhibitors and seven of the 10<br />

(70%) antihyperlipidemic drugs included in the PDR are listed as<br />

having chemosensory side effects. Of the 15 calcium-channel blockers<br />

in the PDR, eight (53%) have been reported to induce taste or smell<br />

disturbances. Moreover, chemosensory dysfunctions have also been<br />

reported with use of angiotensin II antagonists and diuretics. This<br />

abstract also provides in<strong>for</strong>mation on better defining the nature of the<br />

dysfunction, outlines testing strategies and available tests that could be<br />

used to better define the prevalence of the dysfunction, and summarizes<br />

means <strong>for</strong> mitigating such alterations.<br />

This abstract was supported, in part, by Grants PO1 DC 00161, RO1<br />

DC 04278, RO1 DC 02974, and RO1 AG27496 (RLD)<br />

109<br />

412 Poster [ ] Multimodal Integration<br />

REPEATABILITY AND INTERCORRELATIONS OF SENSORY<br />

AND COGNITIVE TESTS IN UNMEDICATED ELDERLY<br />

SUBJECTS<br />

Schiffman S.S. 1, Murray S.G. 1, Watson E.L. 1 1Psychiatry, Duke<br />

University, Durham, NC<br />

The purpose of the study was threefold: 1) to determine if age-related<br />

losses in various sensory functions are correlated, 2) to determine if<br />

sensory acuity is related to cognitive per<strong>for</strong>mance, and 3) to determine<br />

if practice effects occur with repeated testing over a 6 month period.<br />

Elderly individuals between 65 and 85 years of age in good health on<br />

no medications (other than hormone replacement) were tested on the<br />

same battery of cognitive and sensory tests measuring taste, smell,<br />

vision, hearing, and touch. There were a total of 71 subjects (29 male,<br />

42 female) with a mean age of 72.6 years and mean years of education<br />

of 16 years. Subjects were tested at the following time points: baseline,<br />

at 3 weeks, at 2 months, and at 6 months. Data <strong>for</strong> each time point were<br />

collected during two 2-hour sessions per<strong>for</strong>med within 1 week of each<br />

other. The results over the four sessions show considerable variability.<br />

Data analysis indicates rather low correlations between tests of each<br />

sensory modality and with cognition. Repeated testing of elderly<br />

subjects over the 6 month period showed the following trends: 1) selfrated<br />

perception of taste and smell ability decreased, 2) cognitive<br />

per<strong>for</strong>mance improved slightly, and 3) olfactory identification and odor<br />

memory scores decreased slightly. These findings suggest that repeated<br />

testing is necessary to evaluate accurately some sensory and cognitive<br />

functions in the elderly. Supported by a grant from National Institute on<br />

Aging NIA AG 00443.<br />

413 Poster [ ] Multimodal Integration<br />

HEDONIC CONTRAST AND HEDONIC DISCRIMINATION<br />

Allen D. 1, Henley M. 1, Zellner D. 1, Parker S. 2 1Psychology, Montclair<br />

State University, Upper Montclair, NJ; 2Psychology, American<br />

University, Washington, DC<br />

Stimuli are rated as less intense when compared to very intense<br />

context stimuli than when presented alone or with less intense context<br />

stimuli (Frederiksen, 1975), a phenomenon called contrast. In audition,<br />

loud contextual sounds engender a reduction in intensity discrimination,<br />

perhaps by changing the psychophysical loudness function (Parker et<br />

al., 2002). We recently demonstrated contrast with hedonic judgments<br />

(Zellner et al., 2003). Here we look <strong>for</strong> a reduced hedonic<br />

discrimination accompanying hedonic contrast.<br />

In Experiment 1 two groups of subjects rated the pleasantness of<br />

two diluted fruit drinks (Test drinks). Group 1 drank and rated the Test<br />

drinks. Group 2 rated four other full-strength fruit drinks (Context<br />

drinks) be<strong>for</strong>e the Test drinks. Group 2 rated the Test drinks as less<br />

pleasant (M=-22.22) than did Group 1 (M=+8.44), t(32)=2.63, p=.013.<br />

Thus, hedonic contrast occurred.<br />

Experiment 2 investigated whether reduced hedonic discrimination<br />

accompanies that hedonic contrast. Subjects rated how much more they<br />

liked the preferred Test drink than the other one. On the10-point rating<br />

scale, 1 meant “slightly more” and 10 meant “very much more”. Group<br />

1 drank only the diluted Test drinks and rated their difference in<br />

likeability. Group 2 drank and rated them after exposure to the four fullstrength<br />

Context drinks. Group 2 rated the two Test drinks as<br />

hedonically less different (M=1.31) than did Group 1 (M=3.63),<br />

t(30)=2.69, p=.012.<br />

These results demonstrate reduced hedonic discrimination<br />

accompanying hedonic contrast, paralleling phenomena seen in studies<br />

of loudness. Perhaps the psychophysical growth of pleasantness is<br />

malleable.


414 Poster [ ] Multimodal Integration<br />

MEMANTINE AND MECAMYLAMINE ALTER NICOTINE<br />

PERCEPTION IN MAN<br />

Thuerauf N. 1, Lunkenheimer B. 1, Lunkenheimer J. 1, Bleich S. 1,<br />

Schlabeck M. 1, Kornhuber J. 1 1Department of Psychiatry and<br />

Psychotherapy, University of Erlangen-Nürnberg, Erlangen, Germany<br />

NMDA-receptors are present at different regions of the olfactory<br />

system and NMDA-knockout mice were profoundly impaired in<br />

olfactory discrimination. Memantine, a non-competitive, selective<br />

NMDA-antagonist with a voltage-dependent, fast unblocking kinetic<br />

can be used in man. Neuronal nicotinic Acetylcholine receptors<br />

(nAchRs) are present at the nasal mucosa. Mecamylamine, an<br />

Acetylcholine-receptor-antagonist, blocked painful responses to<br />

nicotine following chemical stimulation of the tongue. In the present<br />

study we investigated (1) the influence of Memantine (oral application)<br />

and of (2) Mecamylamine (local application) on nicotine perception<br />

using a placebo controlled double blind (1) parallel group and (2) a<br />

two-fold crossover design. Nicotine perception was studied under<br />

Memantine / placebo (steady state condition, n=15) and under<br />

Mecamaylamine / placebo (following local application, n=15)<br />

following chemical stimulation of the nasal cavity with olfactory and<br />

trigeminal stimulus concentrations. Memantine significantly reduced<br />

the olfactory intensity estimates of S(-)-nicotine and Mecamylamine<br />

significantly reduced trigeminal intensity estimates of nicotine<br />

enantiomers. Thus, Acetylcholine-receptor- and NMDA-antagonists are<br />

effective tools to influence nicotine perception in man acting with<br />

different specificity.<br />

Acknowledgement: Research described in this article was supported<br />

by Philip Morris USA Inc. (Philip Morris External Research Program /<br />

Postdoctoral Fellowship).<br />

415 Poster [ ] Multimodal Integration<br />

MULTIPLE SHORT-TERM EFFECTS OF ENVIRONMENTAL<br />

TOBACCO SMOKE AND PROPIONIC ACID<br />

Suarez J.C. 1, Whitt R. 1, Walker D. 1, Walker J. 1 1Sensory Research<br />

Institute, Florida State University, Tallahassee, FL<br />

As part of a research program to better understand short-term<br />

responses of humans to environmental chemical exposures, we are<br />

measuring multiple responses/impacts from 20 normal (no evidence of<br />

chemosensory deficits or of chemical hypersensitivity) participants (Ps)<br />

in a previously described (ACHEMS 2003) 10-m3 environmental<br />

chamber. Ps are tested in pairs and are given one 100-minute test<br />

session with each of eight stimulus conditions. By varying the<br />

proportion of total chamber flow (fixed at 1000 LPM) derived from an<br />

antechamber in which one or two smokers puff in response to signal<br />

lights, we approximate three target concentrations of environmental<br />

tobacco smoke (ETS): 15, 100 and 800 mg/m3 of ETS-attributable<br />

respirable suspended particulate. As a control, smokers puff on unlit<br />

cigarettes. Four concentrations (0, 1, 10, 15 ppm) of propionic acid<br />

(PA) are generated by varying the proportion of chamber flow passed<br />

over liquid PA held in glass saturators. Except with 15 ppm PA (where<br />

the plateau is held <strong>for</strong> minutes 20-30), concentrations of PA and ETS<br />

rise during minutes 11-20 and then are held through minute 70, when an<br />

exponential decline begins. Sensory/symptom, breathing, eyeblink rate,<br />

psychological state and in<strong>for</strong>mation processing speed responses/impacts<br />

are measured at various times over the course of the session. As a first<br />

step toward an improved understanding of the effects of various<br />

environmental contaminants on short-term responses in humans, our<br />

analyses evaluate stimulus type, concentration and time as determinants<br />

of the magnitude of the change in each endpoint.<br />

Research supported in part by Philip Morris USA Inc.<br />

110<br />

416 Poster [ ] Taste: Umami<br />

COMPARISON OF THE TASTES OF MSG AND L-ALANINE<br />

Taylor-Burds C.C. 1, Westburg A.M. 1, Wifall T.C. 1, Delay E.R. 1<br />

1Neuroscience Program, Regis University, Denver, CO<br />

Monosodium Glutamate (MSG) is an amino acid that occurs<br />

naturally in protein rich foods such as meat and dairy products.<br />

Chaurdhari et al. (1996) showed that MSG activates taste-mGluR4<br />

receptors. Nelson et al. (2002) found that the receptor T1R1/T1R3<br />

responds to many amino acids including MSG, but not to stimuli<br />

perceptually sweet to humans. Alternatively, the T1R2/T1R3 receptor<br />

responds to sugars and artificial sweeteners but not to umami<br />

substances. Nevertheless, conditioned taste aversion (CTA) studies<br />

report that rats perceive similarities in taste between MSG and several<br />

sweet substances when the sodium taste is reduced by amiloride, a<br />

sodium channel blocker (Heyer et al., 2003). Several L-amino acids,<br />

such as L-alanine (ALA), elicit a variety of taste qualities including<br />

“sweet” in humans (Schiffman et al, 1981). Two questions were<br />

addressed in this study: Do ALA and MSG share taste qualities, and if<br />

so can rats distinguish between the two substances? CTA experiments<br />

were done with water-deprived rats conditioned to either ALA or MSG,<br />

and then tested against 3 concentrations of the opposite stimulus, the<br />

conditioned stimulus, and control substances. The results revealed a<br />

strong perceptual similarity between MSG and ALA. To assess the<br />

strength of similarities in these tastes, discrimination experiments were<br />

also conducted. Rats can discriminate between the two substances but<br />

the discrimination is significantly difficult when amiloride or amiloride<br />

+ NaCl is added to control <strong>for</strong> the taste of Na+ in MSG. These results<br />

suggest that MSG and ALA may share similar but not identical afferent<br />

signaling.<br />

417 Poster [ ] Taste: Umami<br />

L-SERINE, GLYCINE, AND MSG TASTES IN RATS:<br />

GENERALIZATION OF CONDITIONED TASTE AVERSION<br />

Mitzelfelt J.D. 1, Westburg A.M. 1, Delay E.R. 1 1Neuroscience Program,<br />

Regis University, Denver, CO<br />

Recent molecular studies suggest that L-amino acids and substances<br />

that elicit an “umami” taste may be detected by T1R1/T1R3<br />

heterodimeric receptors in taste receptor cells whereas “sweet”<br />

substances are detected by T1R2/T1R3 receptors. Monosodium<br />

glutamate (MSG) is a naturally occurring amino acid found in most<br />

protein rich foods such as meats, dairy products, and some vegetables,<br />

and is a prototypical substance that elicits an “umami” taste. In humans,<br />

L-amino acids can elicit quite different tastes (Schiffman et al, 1981).<br />

However, other than preference measures there is little behavioral<br />

research on the taste qualities of amino acids in nonhuman mammalian<br />

species. In this study, conditioned taste aversion (CTA) methods were<br />

used to see if aversions to L-Serine and Glycine, both common and<br />

naturally occurring amino acids, would cross-generalize to L-MSG in<br />

short duration, reactive tests. In contrast to other CTA studies,<br />

amiloride was present to reduce the taste of the sodium ion. A lithium<br />

chloride-induced aversion to 100 mM MSG generalized to L-Serine and<br />

Glycine at concentrations as low as 10 mM when amiloride was present<br />

in all solutions. Similar generalization to MSG was detected when the<br />

rats were conditioned with L-Serine and Glycine. These results suggest<br />

that MSG shares taste qualities with L-Serine and Glycine, possibly<br />

through common afferent signaling mechanisms. Supported by NIH<br />

R15DC005962


418 Poster [ ] Taste: Umami<br />

POLYMORPHISMS OF KNOWN MONOSODIUM<br />

GLUTAMATE (MSG) TASTE RECEPTORS AND BIMODAL<br />

DISTRIBUTION OF SENSITIVITY TO GLUTAMATE SAVORY<br />

TASTE<br />

Alarcon S.M. 1, Mascioli K.J. 1, Reed D. 1, Keast R. 1, Tharp C. 1, Lui S. 1,<br />

Ahmed O. 1, Breslin P. 1 1Monell Chemical Senses Center, Philadelphia,<br />

PA<br />

Monosodium glutamate (MSG) elicits a unique taste quality, often<br />

labeled savory, brothy, or umami. This quality is thought to underlie the<br />

protein and amino acid tastes, which convey a preferred flavor in many<br />

foods. But a small subset of the population (~6%) is insensitive to this<br />

taste quality. Subjects were screened <strong>for</strong> their ability to distinguish 29<br />

mM NaCl vs 29 mM MSG in a <strong>for</strong>ced-choice triangle test. The few<br />

subjects who could not discriminate these stimuli also demonstrated an<br />

insensitivity to glutamate on a concentration-intensity scaling task.<br />

Furthermore, the synergy of savory taste that occurs <strong>for</strong> most subjects<br />

when MSG is mixed with 5' ribonuclotides, such as GMP or IMP, is<br />

lacking in selected individuals and can occur despite normal sensitivity<br />

to MSG savoriness. There<strong>for</strong>e, there are semi-independent mechanisms<br />

underlying sensitivity to MSG and its interactions with IMP and GMP.<br />

At present, there are two putative savory taste metabotropic receptors<br />

identified: human taste GRM4 [6p21.3] and the receptor dimer human<br />

TAS1R1-TAS1R3 [1p36.23; 1p36.33]. Taste GRM4 (an orally located<br />

splice variant of the CNS GRM4, the group lll inhibitory glutamate<br />

receptor) was sequenced and several identified exonic polymorphisms<br />

were not correlated with savory taster status. The common<br />

polymorphism in TAS1R3 (A5T) was also not correlated with savory<br />

taster status. TAS1R1 contains eight less common polymorphisms,<br />

including a nonsense mutation, comprising several haplotypes in our<br />

subjects. TAS1R1 also has four splice variants. These studies will help<br />

identify the receptors responsible <strong>for</strong> this quality of taste in humans.<br />

Research was funded by DC02995 (breslin@monell.org).<br />

419 Poster [ ] Taste: Umami<br />

GLUTAMATE- AND INOSINE-INDUCED CALCIUM<br />

RESPONSES IN MOUSE TASTE RECEPTOR CELLS<br />

Maruyama Y. 1, Chaudhari N. 2, Roper S.D. 2 1Physiology & Biophysics,<br />

University of Miami School of Medicine, Miami, FL; 2Physiology &<br />

Biophysics and Program in Neuroscience, University of Miami, Miami,<br />

FL<br />

L-glutamate (L-Glu) is a naturally occurring amino acid in proteinrich<br />

food and elicits umami taste. Umami taste is synergistically<br />

enhanced by nucleotides, including inosine 5'-monophosphate (IMP).<br />

Hypotheses proposed to explain nucleotide synergy include: 1) separate<br />

receptors <strong>for</strong> L-Glu and nucleotide are expressed in distinct cells and<br />

integration occurs at a subsequent stage, e.g. postsynaptically; 2) each<br />

receptor is expressed in the same taste receptor cells; 3) L-Glu and<br />

nucleotide interact with the same receptor. To examine whether the<br />

same taste cells respond to L-Glu and IMP, we measured intracellular<br />

Ca2+ responses to L-Glu, IMP and a mixture of both tastants in<br />

gustatory sensory cells loaded with Calcium Green-1 Dextran and<br />

imaged in slices of circumvallate papillae with confocal microscopy.<br />

Monopotassium glutamate (MPG, 500 mM) alone, as well as IMP (1<br />

mM) alone, applied focally to taste pores, induced transient [Ca2+ ] i<br />

responses in some taste receptor cells. To date, MPG-responsive cells<br />

have also responded to IMP, suggesting that glutamate and nucleotides<br />

act on the same receptor. Alternatively, the cognate receptors are<br />

expressed in the same cell and the receptor(s) do not obligatorily<br />

require a mixture of IMP and glutamate. Mixtures of MPG and IMP<br />

elicited ~3-fold larger responses than the summation of individual<br />

responses to MPG and IMP alone. These data indicate that nucleotide<br />

synergy of glutamate taste is detectable at the level of Ca2+ signals in<br />

individual taste cells. Supported by NIH/NIDCD grant DC03013.<br />

111<br />

420 Poster [ ] Taste: Umami<br />

BEHAVIORAL TASTE RESPONSES TO MONOSODIUM<br />

GLUTAMATE AND SODIUM CHLORIDE IN FOUR SPECIES<br />

OF NONHUMAN PRIMATES<br />

Hernandez Salazar L. 1, Laska M. 2 1Instituto de Neuro-Etologia,<br />

Universidad Veracruzana, Xalapa, Mexico; 2Department of Medical<br />

Psychology, University of Munich, Munich, Germany<br />

Taste responses of six squirrel monkeys, five pigtail macaques, four<br />

olive baboons and four spider monkeys to monosodium glutamate<br />

(MSG) and to sodium chloride were assessed in two-bottle preference<br />

tests of brief duration (2 min). When given the choice between tap<br />

water and defined concentrations of the two tastants dissolved in tap<br />

water, the animals were found to significantly discriminate<br />

concentrations of MSG as low as 2 mM (spider monkeys and olive<br />

baboons), 50 mM (pigtail macaques) and 300 mM (squirrel monkeys)<br />

from the solvent. With sodium chloride, taste preference thresholds<br />

were found to be 1 mM (spider monkeys), 20 mM (pigtail macaques),<br />

50 mM (olive baboons), and 200 mM (squirrel monkeys), respectively.<br />

Across-species comparisons of the degree of preference <strong>for</strong> MSG and<br />

sodium chloride displayed by the four primate species showed the same<br />

order of spider monkeys > olive baboons > pigtail macaques > squirrel<br />

monkeys. When presented with equimolar concentrations of different<br />

tastants, all four species preferred sucrose as well as a mixture of<br />

sucrose and sodium chloride over MSG, and – at least at one<br />

concentration – they preferred MSG over sodium chloride. The results<br />

support the assumption that the taste responses of the four primate<br />

species to MSG and sodium chloride might reflect an evolutionary<br />

adaptation to their respective dietary habits.<br />

421 Poster [ ] Human Olfactory Per<strong>for</strong>mance<br />

SOURCE MEMORY FOR ODORS AND OBJECTS IN<br />

CHILDREN AND YOUNG ADULTS<br />

Pirogovsky E. 1, Gilbert P. 2, Addie S. 1, Stull K. 1, Ricardo T. 1, Viera E. 1,<br />

Murphy C. 1 1Psychology, San Diego State University, San Diego, CA;<br />

2School of Medicine, University of Cali<strong>for</strong>nia San Diego, San Diego,<br />

CA<br />

Recall and recognition memory <strong>for</strong> odors is poorer in children than in<br />

adolescents. In addition, children per<strong>for</strong>m worse than young adults in<br />

source memory tasks using visual and auditory stimuli. However,<br />

source memory <strong>for</strong> odor stimuli has not been examined in children. The<br />

present study investigated source and item memory <strong>for</strong> odors and<br />

objects in children (7-10 yrs) and young adults (18-24 yrs). During the<br />

study phase, 1 male and 1 female experimenter (sources) randomly<br />

presented either 16 odors or 16 objects to the participant. Presentation<br />

alternated between sources so that each source presented 8 stimuli.<br />

Once the16 stimuli were presented, the sources exited and a third<br />

experimenter began the test phase. To assess item recognition memory,<br />

a stimulus from the study phase and a novel stimulus were presented to<br />

the participant who was asked to choose the stimulus presented in the<br />

study phase. Source memory was assessed with the 8 stimuli not used in<br />

the item memory task. The experimenter presented a stimulus and asked<br />

whether the male or female experimenter had previously presented the<br />

stimulus. Results indicate that <strong>for</strong> objects, source and item memory<br />

were similar <strong>for</strong> children and adults. For odors, item memory was<br />

similar <strong>for</strong> the two groups; however, source memory <strong>for</strong> odors was<br />

significantly poorer in children than in young adults. It has been<br />

suggested that the frontal lobes play a critical role in source memory<br />

and odor memory and that this brain region continues to develop into<br />

adolescence. Children´s poor per<strong>for</strong>mance on the source memory task<br />

<strong>for</strong> odors may be due in part to the immaturity of the frontal lobes.<br />

Supported by NIH grant #AG04085 to CM


422 Poster [ ] Human Olfactory Per<strong>for</strong>mance<br />

AGE-RELATED CHANGES IN SOURCE AND ITEM MEMORY<br />

FOR ODORS AND OBJECTS<br />

Gilbert P.E. 1, Ferdon S. 2, Pirogovsky E. 2, Moreland M. 2, Owens A. 2,<br />

Wilkes A. 2, Murphy C. 2 1Head and Neck Surgery, University of<br />

Cali<strong>for</strong>nia, San Diego, San Diego, CA; 2Psychology, San Diego State<br />

University, San Diego, CA<br />

Source memory is more affected by aging than item memory.<br />

However, little research has investigated age-related changes in source<br />

and item memory <strong>for</strong> olfactory stimuli. The present study compared<br />

source and item memory <strong>for</strong> odors and objects in healthy elderly and<br />

young adults. During the study phase, either 16 odors or 16 objects<br />

were randomly presented by a male and a female experimenter<br />

(sources) to the participant. Presentation alternated between sources so<br />

that 8 stimuli were presented by each source. Once the stimuli were<br />

presented, the test phase was conducted by a third experimenter. To<br />

assess source memory, a stimulus from the study phase was presented<br />

to the participant who was asked to indicate whether the stimulus was<br />

presented by the male or female experimenter. To assess item<br />

recognition memory, a stimulus from the study phase and a novel<br />

stimulus were presented and the participant was asked to indicate which<br />

stimulus was presented during the study phase. Source and item<br />

memory <strong>for</strong> odors and objects were similar in young adults. In elderly<br />

adults, source memory was impaired relative to item memory <strong>for</strong> odors<br />

and objects. However, the impairment in source memory relative to<br />

item memory <strong>for</strong> odors was significantly greater than <strong>for</strong> objects. The<br />

data suggest that source memory <strong>for</strong> olfactory stimuli may be<br />

particularly sensitive to age-related changes in the brain. Supported by<br />

NIH grant AG04085 to Claire Murphy and training grant DC00032 to<br />

Terence M. Davidson.<br />

423 Poster [ ] Human Olfactory Per<strong>for</strong>mance<br />

AGE DEPENDENT ODOUR MEMORY AND ODOUR<br />

IDENTIFICATION: DIFFERENTIAL EFFECTS OF GENDER<br />

Møller P. 1, Wulff C. 1 1Royal Veterinary and Agricultural University,<br />

Frederiksberg, Denmark<br />

Objective: To investigate the decline of odour memory and odour<br />

identification with age.<br />

Method: Ten different well-known food odours were presented in a<br />

4AFC identification task to 279 subjects (78 males) younger and 229<br />

subjects (73 males) older than 50 years of age. Half of the subjects in<br />

each age group were cued to remember three of the ten odours, whereas<br />

the other half were not in<strong>for</strong>med about the following memory task.<br />

Memory <strong>for</strong> the same target odours in the two groups were tested with a<br />

3AFC procedure, where each target odour was to be detected among<br />

two distractor odours.<br />

Results: Mann-Whitney tests of overall memory find no significant<br />

differences between males and females in the groups below and above<br />

50 years of age. If subjects are divided by the age of 60 (202 Ss older<br />

than 60, 68 males) elderly women outper<strong>for</strong>m elderly men significantly<br />

(p


426 Poster [ ] Human Olfactory Per<strong>for</strong>mance<br />

ODORS AND MEMORIES: THE PROUST PHENOMENON<br />

REVISITED<br />

Hudson J.A. 1, Wilson P. 2, Freyberg R. 1, Haviland-Jones J. 1<br />

1Psychology, Rutgers, The State University of New Jersey, Piscataway,<br />

NJ; 2LaSalle University., Philadelphia, PA<br />

The Proust phenomenon refers to the common belief that odor cues<br />

elicit vivid and emotional childhood memories. To test the validity of<br />

this phenomenon, we investigated the effects of odors alone, in the<br />

absence of semantic or visual cues, on autobiographic memory. 110<br />

participants (N=55 men, 55 women) were exposed to either a control<br />

(no odor) or an ambient odor (one of 3 childhood, 3 environmental or 3<br />

floral odors) while writing childhood and recent memories. ANOVAS<br />

compared emotion words, sensory references, and emphatic language<br />

across odor categories. We found no evidence <strong>for</strong> a Proust<br />

phenomenon. There was no significant difference in emotion, sensory<br />

qualities or emphatic language in childhood memories as a function of<br />

odor condition. When odor effects were obtained, they were effects of<br />

floral odors on recent memories, not childhood memories (all reported<br />

effects, p < .05). Recent memories included more positive emotion<br />

words in the floral odor condition as compared to all other conditions.<br />

Women reported less negative emotion in the floral condition <strong>for</strong> recent<br />

memories and men reported more sensory qualities in the floral odor<br />

condition. Thus, (a) childhood odors had no special effect on childhood<br />

memories (b) there was no overall odor effect on childhood memories;<br />

and (c) odors did not affect childhood memories more than recent<br />

memories. In contrast, recent memories were more affected by odor<br />

than childhood memories.<br />

Funded by International Flavors & Fragrances, Inc.<br />

427 Poster [ ] Human Olfactory Per<strong>for</strong>mance<br />

ODOR CATEGORIZATION IN THE ABSENCE OR IN THE<br />

PRESENCE OF ODOR NAMES<br />

David S. 1, Rouby C. 2, Bensafi M. 2, Barkat S. 2 1CNRS UMR 7114 –<br />

Modyco et Université Paris 10, Paris, France; 2CNRS UMR 5020 et<br />

Universite Lyon1, Lyon, France<br />

In order to study the top-down influence of names on odor<br />

categorization, we compared situations allowing perceptual<br />

categorization which is supposed to be mainly driven by stimulus<br />

properties, with semantic categorization supposed to be driven by our<br />

knowledge. Two groups of 20 subjects were asked to <strong>for</strong>m categories<br />

among 16 odors according to their similarity : group 1 was provided<br />

odor vials without any in<strong>for</strong>mation on the name, group 2 was provided<br />

the same odor vials labeled with a name. A third group categorized only<br />

the labels, without olfactory stimuli, so relying on their memory to<br />

judge similarity. After the categorization task, subjects evaluated the 16<br />

odors again on intensity, pleasantness, familiarity and typicality. We<br />

stated the hypothesis that when judging real odors, the main dimension<br />

<strong>for</strong> odor categorization would be the hedonic valence, and that valence<br />

would no longer dominate when categorizing from memory. The results<br />

confirmed the hypothesis : hedonic ratings were correlated with the<br />

main dimension <strong>for</strong> groups 1 and 2, not <strong>for</strong> group 3. Results also show<br />

that the presence of names reduces the number of categories and that<br />

these categories tend to be hierarchically organized, which is not the<br />

case <strong>for</strong> the group 1.<br />

113<br />

428 Poster [ ] Human Olfactory Per<strong>for</strong>mance<br />

CAN YOUR NOSE SHINE AN ATTENTIONAL SPOTLIGHT?<br />

Porter J.A. 1, Zelano C. 1, Mainland J. 1, Johnson B. 1, Bremner E. 1, Khan<br />

R.M. 1, Bensafi M. 1, Sobel N. 1 1Biophysics, University of Cali<strong>for</strong>nia,<br />

Berkeley, Berkeley, CA<br />

Directed spatial attention in vision and audition enables heightened<br />

sensory acuity <strong>for</strong> events at known spatial locations. We asked whether<br />

a similar phenomenon exists in chemical sensing. Initially, we asked if<br />

humans have access to spatial in<strong>for</strong>mation through chemosensation. We<br />

tested 33 subjects on a 16 trial left vs. right localization task. The<br />

odorants rose and strawberry (diluted 1:4 into proprionic acid) were<br />

generated by an air dilution olfactometer, and delivered via a divided<br />

nasal mask with separate left and right entry ports. As a group, subjects<br />

were slightly but significantly above chance in their ability to spatially<br />

localize the odorants (accuracy = 57.8%±2.7% t=2.841, p


430 Poster [ ] Human Olfactory Per<strong>for</strong>mance<br />

DISCRIMINATING DEUTERATED FROM UNDEUTERATED<br />

ACETOPHENONE: COMPARING HUMANS AND A DOG<br />

Zelano C. 1, Rubenstein N. 1, Mainland J. 1, Porter J. 1, Bremner E. 1,<br />

Johnson B. 1, Khan R. 1, Sobel N. 1 1Biopyhsics, University of Cali<strong>for</strong>nia,<br />

Berkeley, Berkeley, CA<br />

To determine if deuterated compounds can be distinguished from<br />

their undeuterated counterparts through smell, we set out to replicate a<br />

previous report: 31 subjects per<strong>for</strong>med a one-trial same/different<br />

discrimination on two jars containing deuterated (D) and undeuterated<br />

(H) acetophenone. 24 subjects correctly chose different. However,<br />

when per<strong>for</strong>ming the same task using two identical samples of H, 23<br />

subjects erroneously chose different. Thus, a bias existed toward<br />

answering different. To avoid the bias, we used an olfactometer to test<br />

38 subjects in a 32 trial same/different <strong>for</strong>ced choice design. Mean<br />

accuracy was not different from chance (52.5% ± 3.5, p>.1). To<br />

eliminate possible habituation effects, we repeated the task in 14<br />

subjects with only 8 trials. Mean accuracy was not different from<br />

chance (60% ± 23, p>.1). To test a different paradigm, an additional 10<br />

subjects per<strong>for</strong>med a three-alternative <strong>for</strong>ced choice identification task.<br />

Mean accuracy was not different from chance (38.5%±22, p>.1).<br />

Humans thus far unsuccessful, we set out to see if a dog could<br />

discriminate H from D. A German Shepard was trained to recognize H.<br />

He then did 7 trials of a three-alternative <strong>for</strong>ced-choice identification<br />

task at 100% accuracy (p


Abou, Elizabeth, 309<br />

Abraham, Michael H., 67<br />

Abrell, Leif, 267<br />

Aburatani, Hiroyuki, 199<br />

Acevedo, Humberto P, 290<br />

Ache, Barry W, 47, 216<br />

Adams, Chris, 63<br />

Adams, Julie, 246<br />

Addie, Servin, 421<br />

Agerman, Karin, 205<br />

Aggio, Juan F., 216<br />

Ahmed, Osama, 418<br />

Akins, Michael, 222, 256<br />

Alarcon, Suzanne M, 298, 418<br />

Albrecht, Gudrun, 331<br />

Alex, James, 306<br />

Alimohammadi, Hessamedin, 172, 173<br />

Allen, Dawn, 413<br />

Allgood, Sallie, 171<br />

Alony, Ronny, 127<br />

Alvarez-Buylla, Arturo, 1<br />

Amendola, Diedra, 154<br />

Amer, Ahmed, 63<br />

Ang, Lay-Hong, 9<br />

Ansari, Rahman, 104<br />

Apkarian, Apkar Vania, 26<br />

Arbuckle, Wesley, 157, 284<br />

Arellano, Julie, 45<br />

Armstrong, Cheryl L. H., 30<br />

Arriola, Aline, 378<br />

Ashkenazi, Amir, 302<br />

Atarot, Tal, 127<br />

Atema, Jelle, 237, 265, 406<br />

Atkinson, Rachel, 371<br />

AtKisson, Mary S., 125<br />

Axel, Richard, 72<br />

Bachmanov, Alexander, 90, 296<br />

Bai, Li, 204<br />

Bailey, Michele L., 115<br />

Bailie, Jason M, 118, 120, 121<br />

Baird, John P, 316<br />

Baker, Harriet, 381<br />

Balderston, Catherine C, 332<br />

Baldwin, David, 376<br />

Baquero, Arian, 181<br />

Barbara, Cerf, 251<br />

Barbour, Jon, 130, 358<br />

Bargmann, Cornelia I., 70<br />

Barkat, Samy, 427<br />

Barnard, Joan, 245<br />

Barrows, Jennell, 83<br />

Bartoshuk, Linda M., 21, 215, 304, 305, 306, 366<br />

Battey, James F., 211<br />

Bayevsky, Alex, 319<br />

Beauchamp, Gary K., 13, 58, 296<br />

Bedner, Peter, 362<br />

Behan, John, 104<br />

115<br />

Beites, Crestina, 11<br />

Belanger, Andrea, 157<br />

Belanger, Rachelle M., 52, 236<br />

Bell, Wade E., 108<br />

Ben-Asher, Edna, 127<br />

Benard, Lumie, 182, 295<br />

Bensafi, Moustafa, 103, 310, 427, 428<br />

Benton, Richard, 357<br />

Bergman, Daniel A., 52, 234, 235<br />

Berlin, RoseAnn, 381<br />

Bernd, Bufe, 191, 433<br />

Berndtsson, Jonas, 164<br />

Best, Aaron Robert, 99<br />

Bhandawat, Vikas, 51<br />

Bhatt-Mackin, Seamus Michael, 26<br />

Bigbee, John W, 188<br />

Bilder, Robert M, 331<br />

Blake, Camille, 96<br />

Blakemore, Laura J., 396<br />

Bleich, Stefan, 414<br />

Blizard, David A., 369<br />

Blonde, Ginger D., 79<br />

Bobkov, Yuriy V, 47<br />

Bohbot, Jonathan, 136, 268<br />

Bomsztyk, Mayan, 208<br />

Bönigk, Wolfgang, 50<br />

Bose, Soma C., 286<br />

Böttger, Bärbel, 172<br />

Boughter, John D., 186, 318, 368<br />

Boughter, Jr, John D., 213<br />

Boyle, Julie A., 101, 102<br />

Bradley, Jonathan, 50, 279<br />

Bradley, Robert M., 92<br />

Brand, Joseph, 183, 328, 330<br />

Brann, Jessica H., 35<br />

Brasser, Susan M, 93, 209<br />

Bremner, Elizabeth, 310, 428, 430, 432<br />

Breslin, Paul A.S., 23, 24, 25, 166, 191, 298, 418, 433<br />

Breza, Joseph M., 85<br />

Brown, Heather J., 108<br />

Brunjes, Peter C., 8<br />

Bryant, Bruce, 170<br />

Buck, Linda, 311<br />

Buntinas, Linas, 292<br />

Burd, Gail, 140<br />

Burgess, John R., 31<br />

Burks, Catherine A, 74<br />

Buschmann-Maiworm, Regina, 123<br />

Byrd, Christine A., 221<br />

Cain, William S., 67<br />

Calder, Andrew J, 26<br />

Calhoun-Haney, Rose, 251<br />

Calof, Anne L., 11<br />

Campbell, Hannah, 17<br />

Cao, Jie, 328, 330<br />

Caprio, John, 97<br />

Caramanos, Zografos, 68


Carlson, John R, 129, 131<br />

Carlsson, Mikael, 3<br />

Carlton, Michelle, 62<br />

Carr, Virginia McM, 206<br />

Carson, Christine, 351<br />

Case, Gilbert R., 96, 275<br />

Cerf-Ducastel, Barbara, 303, 309<br />

Chalé, Angela, 31<br />

Chandran, Suchismita, 106, 124, 283<br />

Chapo, Audrey, 21, 306<br />

Chaudhari, Nirupa, 80, 419<br />

Chen, Denise, 162, 163<br />

Chen, Margaret, 309<br />

Chen, Ping, 32, 33, 278<br />

Chen, Wei R., 395<br />

Chen, Wen, 63<br />

Chess, Andrew, 145<br />

Chi, Qiuyi, 128<br />

Chin, Joanna, 349<br />

Chopra, Anita, 65<br />

Choudhury, Eric, 122<br />

Christ-Hazelhof, Elly, 15<br />

Christensen, Thomas, 138, 230, 403, 407<br />

Christiansson, Sven Åke, 164<br />

Christina, Kuhn, 191<br />

Christy, Robert C., 87, 186<br />

Cinelli, Angel, 33, 278<br />

Clapp, Tod, 184<br />

Clark, Larry, 271<br />

Cleland, Thomas, 60, 231, 232<br />

Clevenger, Amy C., 57<br />

Clyburn, Virginia L, 28<br />

Coelho, Daniel, 306<br />

Cohen, Lawrence B., 7<br />

Colbert, Connie L., 78, 319<br />

Coleman, Elaine S, 386<br />

Colley, Beverly Shelley, 393<br />

Collmann, Chad, 3, 44<br />

Cometto-Muniz, Jorge Enrique, 67<br />

Conley, David B, 337, 348<br />

Connelly, Timothy, 335<br />

Contreras, Robert J., 27, 85<br />

Cook, Michelle, 259<br />

Cooney, Janine, 42<br />

Corkum, Lynda, 157<br />

Cornelia, Hummel, 109<br />

Corotto, Frank Stuart, 404<br />

Costanzo, Richard M., 41<br />

Coureaud, Gerard, 312<br />

Cowart, Beverly J., 119, 342<br />

Crasto, Chiquito J, 363<br />

Crocker, Candice, 11<br />

Cunningham, Anne M., 144<br />

Curran, Maryanne, 58<br />

Curtis, Kathleen S, 27, 85<br />

Cusick, Matthew, 283<br />

Dacks, Andrew, 407<br />

116<br />

Dalton, Pamela, 116, 117, 166, 167, 176, 338, 342<br />

Dalton, Wang, 33<br />

Daly, Kevin, 216, 219, 220<br />

Daly, Mark, 145<br />

Damak, Sami, 189, 370<br />

Damann, Nils, 175<br />

Daniel, Peter Carter, 53<br />

Daniels, Yasmine, 38<br />

Danilov, Yuri, 370<br />

Danilova, Vicktoria, 370<br />

Dankulich, Luba, 291<br />

Darby-King, Andrea, 61<br />

Darmohusodo, Vincent, 361<br />

Davenport, Rachel A., 317<br />

David, Sophie, 427<br />

Davidson, Andrew, 304, 366<br />

Davis, Michael, 59<br />

Davison, Ian G., 224<br />

Day, Kristen M., 390<br />

De Wijk, Rene, 22, 409<br />

DeFazio, Richard Anthony, 81<br />

Delay, Eugene R., 416, 417<br />

Delay, Rona, 155, 280, 281, 283<br />

Delwiche, Jeannine, 16<br />

DeMarchis, Silvia, 257<br />

DenBleyker, Megan J., 294<br />

Dennis, John Carroll, 386<br />

Derby, Charles D., 238, 239, 247, 269, 270, 388, 389<br />

DeSimone, John A., 75<br />

Desimone, John A., 76, 82, 188<br />

Devanand, Devangere P., 193<br />

Di Lorenzo, Patricia, 88, 91<br />

Diamond, Jeanmarie, 166<br />

Dinehart, Mary E., 215<br />

Distel, Hans, 378<br />

Dittman, Andrew, 376<br />

Djordjevic, Jelena, 102<br />

Dodds, Tyler, 256, 349<br />

Dolensek, Jurij, 153<br />

Dotson, Cedrick D., 210<br />

Doty, Richard L., 122, 192, 195, 341, 335, 379, 411<br />

Dowling, Ryan S, 84<br />

Drago, Joan, 165<br />

Drayna, D., 191, 367, 433<br />

Duffy, Valerie B., 19, 21, 215, 304, 306, 366<br />

Dulchenko, Natalia, 271<br />

Dvoriantchikov, Guennadiy, 80<br />

Eisthen, Heather L., 156<br />

El Sharaby, Ashraf, 200<br />

Ennis, Matthew, 397, 398, 399<br />

Erdelyi, F., 397<br />

Erisir, Alev, 321<br />

Erland, Susanne, 374<br />

Ern<strong>for</strong>s, Patrik, 205<br />

Eslinger, Paul J., 104<br />

Essick, Greg, 65<br />

Eylam, Shachar, 77


Fadool, Debra, 34, 35, 64, 392, 393<br />

Farahbod, Haleh, 228<br />

Farbman, Albert I., 206, 345<br />

Farmer, Jennifer M, 335<br />

Farmer-George, Megan, 69<br />

Fasolo, Aldo, 257<br />

Fatsis, Stefan, 116<br />

Fee, Michale, 225<br />

Felber, Stefan, 331<br />

Feld, I, 163<br />

Feldman, George, 178<br />

Feldman, Roy S., 330<br />

Feldmesser, Ester, 127<br />

Ferdon, Sally, 251, 422<br />

Fernandez, Kenny, 136<br />

Fernando, Shahan, 44<br />

Findley, Leslie, 333<br />

Fine, Jared M., 264<br />

Finger, Thomas E., 83, 86, 148, 172<br />

Firby, Ashley E., 284<br />

Flecke, Christian, 375<br />

Fleischhacker, W Wolfgang, 331<br />

Fletcher, Max, 226<br />

Floriano, Wely B., 126, 367<br />

Forestell, Catherine A., 159<br />

Formaker, Bradley K, 84<br />

Foster, James D., 383<br />

Frank, Marion E., 84, 212, 302, 369<br />

Frank, Robert A., 118, 120, 121<br />

Frasnelli, Johannes, 109, 174, 177, 277, 339<br />

Freyberg, Robin, 425, 426<br />

Frings, Stephan, 50<br />

Fukami, Hideyuki, 92<br />

Fukuda, Nanaho, 360<br />

Fuller, Cynthia L., 221<br />

Fung, France W., 351<br />

Fusetani, Nobuhiro, 240<br />

Fushiki, Tohru, 29<br />

Fussnecker, Brendon L., 62<br />

Futschik, Thomas, 177<br />

Galindo-Cuspinera, Veronica, 24<br />

Gallagher, Michela, 308<br />

Garcea, Mircea, 78, 79, 319<br />

Gelperin, Alan, 225, 233<br />

Gensch, Thomas, 50<br />

Gent, Janneane F., 302<br />

George, Pravin, 300<br />

Gerber, Johannes C, 277<br />

Gerlach, Gabriele, 265, 373<br />

Gesteland, Robert C, 118, 120, 121<br />

Getchell, M. L., 346, 347<br />

Getchell, T. V., 346, 347<br />

Ghosh, Shobha, 188<br />

Gibson, Nicholas J., 137, 139<br />

Gilad, Yoav, 127<br />

Gilbert, Paul E, 421, 422<br />

117<br />

Gilbertson, Timothy A, 69, 74, 179, 180, 181<br />

Gin, Christopher, 141<br />

Gisselmann, Günter, 130<br />

Glendinning, John I, 189, 208<br />

Goddard, William A., 126, 367<br />

Goins, Michael, 307, 343<br />

Goldman, Aaron, 131<br />

Goldman, Daniel J., 221<br />

Gomez, George, 154, 291<br />

Gorman, Rachael, 280<br />

Gould, Fred L., 55<br />

Gould, Michele, 338<br />

Graham, Kristin, 112<br />

Green, Barry, 300<br />

Green, Elgin, 402<br />

Green, Mariah H., 108<br />

Greenaway, Peter, 374<br />

Greenwood, David, 42, 314<br />

Greer, Charles A., 222, 256, 349<br />

Griff, Edwin R., 400, 401<br />

Groot, A, 55<br />

Grozinger, Christina M., 372<br />

Guenter, Jeremy, 180<br />

Guerenstein, Pablo G., 267<br />

Guest, Steve, 65<br />

Gulbransen, Brian, 172<br />

Gunnarson, Amy, 135<br />

Gur, Raquel E, 332<br />

Gutierrez, Ranier, 94<br />

Haas, Lori, 309<br />

Haase, Lori B., 303<br />

Hagelin, Julie, 241, 242, 243<br />

Hahn, Chang Gyu, 384<br />

Halbich, Wolfgang, 218<br />

Hall, Spencer, 126<br />

Hallem, Elissa A., 129<br />

Hallock, Robert M., 88, 91<br />

Halpern, Bruce Peter, 429<br />

Halpern, Mimi, 32, 33, 38, 40, 273, 274, 278<br />

Hamilton, K. A., 397<br />

Hänel, Tobias, 174<br />

Hansen, Anne, 148<br />

Hansen, Dane R, 69, 74, 179<br />

Hansson, Bill, 3, 374<br />

Hao, Yan-Peng, 391<br />

Hardin, Debra, 286<br />

Harley, Carolyn W, 61<br />

Harteneck, Christian, 362<br />

Haskins, Mark, 291<br />

Hastings, Lloyd, 110<br />

Hatt, Hanns, 130, 168, 175, 358<br />

Haviland-Jones, Jeannette, 425, 426<br />

Hawkes, Christopher H., 196, 333<br />

Haxel, Boris R., 336<br />

Hayar, Abdallah, 397, 399<br />

Hayashi, Hisaki, 11<br />

Hayes, John, 19, 21


Heck, Gerard L, 75, 76, 82, 178, 188<br />

Hegg, Colleen, 254, 285<br />

Heidema, Johannes, 15<br />

Heilmann, Stefan, 130, 174, 339<br />

Heilmann, Stefan K., 408<br />

Heinbockel, Thomas, 397, 398, 399<br />

Heldt, Scott, 59<br />

Hellekant, Goran, 370<br />

Henion, Timothy, 377<br />

Henkel, Marta, 123<br />

Henley, Monique, 413<br />

Hennet, Thierry, 377<br />

Hernandez Salazar, Laura Teresa, 420<br />

Hettinger, Thomas P., 212, 369<br />

Higuchi, Makoto, 379<br />

Hildebrand, John G., 138, 230, 267, 407<br />

Hile, Arla, 243<br />

Hill, David L., 321, 322<br />

Hillier, Kirk N, 55, 217<br />

Hing, Huey K., 9<br />

Hingco, Edna E., 350<br />

Hino, Akihiro, 12<br />

Hinterhuber, Hartmann, 331<br />

Hirano, Ken-ichi, 29<br />

Hirsch, Alan R., 115, 344<br />

Ho, Michael G., 129<br />

Hoe, Lori, 386<br />

Hoffman, S M, 118<br />

Horner, Amy J., 238<br />

Horning, K A, 118<br />

Hornung, David, 56<br />

Houpt, Thomas A., 317<br />

Hsu, Jessie, 352<br />

Huang, Liquan, 183, 328<br />

HUANG, YI-JEN J, 329<br />

Hudry, Julie, 424<br />

Hudson, Judith A., 425, 426<br />

Hudson, Robyn, 378<br />

Huesa, Gema, 86<br />

Huettenbrink, Karl-Bernd, 177<br />

Hultine, Stacy, 58<br />

Hummel, Patrick A., 126<br />

Hummel, Thomas, 109, 130, 165, 174, 177, 277, 339, 340, 408<br />

Huque, Taufiqul, 330<br />

Hurst, Jane, 313<br />

Hutchinson, Ian, 253<br />

Hüttenbrink, Karl-Bernd, 339. 340<br />

Hyder, Fahmeed, 4<br />

Ichikawa, Masumi, 37<br />

Illig, Kurt R., 8<br />

Inoue, Masashi, 296<br />

Iwatsuki, Ken, 199<br />

Iwema, Carrie, 256, 349<br />

Jacobson, Aaron, 303<br />

Jafek, Bruce W., 114<br />

Jahng, Jeong Won, 320<br />

118<br />

Jakob, Ingrid, 158<br />

Jang, Woochan, 143<br />

Jean, Gotman, 334<br />

Jensen, Dwayne, 42<br />

Jeon, S-Y, 297<br />

Ji, Qingzhou, 182<br />

Jia, Changping, 40, 274<br />

Jia, Fan, 395<br />

Jiang, Peihua, 295<br />

Jinks, Anthony, 253<br />

Johns, Malcolm, 388<br />

Johnson, Brad, 103, 310, 364, 428, 430, 432<br />

Johnson, Brett Alan, 227, 228, 350<br />

Johnson, Paul, 269, 270<br />

Johnston, Megan Elizabeth, 404<br />

Johnston, Robert E., 315<br />

Jones, David, 371<br />

Jones, Seth V., 59<br />

Jones-Gotman, Marilyn K., 68, 101, 102, 334, 424<br />

Jordan, Melissa, 42<br />

Jorde, Lynn, 367<br />

Judd, Robert L, 386<br />

Julie, Hudry, 334<br />

Jung, Yewah, 154<br />

Kajii, Yuka, 29<br />

Kamio, Michiya, 239<br />

Kashyap, Manoj K, 202<br />

Kataoka, Hiroshi, 360<br />

Katdare, Ameeta, 163<br />

Katsuie, Yasutomi, 369<br />

Katsukawa, Hideo, 293<br />

Katz, Lawrence Charles, 5, 224<br />

Kauer, John S., 125, 289, 365<br />

Kaupp, Benjamin, 50<br />

Kawada, Teruo, 29<br />

Kawai, Takayuki, 29<br />

Kawauchi, Shimako, 11<br />

Kay, Leslie M, 380<br />

Keast, Russell SJ, 25, 418<br />

Keiser, Meagan, 107<br />

Keller, Andreas, 250<br />

Keller, Gerald S, 207<br />

Kemmler, Georg, 331<br />

Kemmotsu, Nobuko, 431<br />

Kennedy, Janice M., 13, 14, 214<br />

Kent, Lauren B., 354<br />

Kern, Robert C, 337, 348<br />

Kerr, Kara-Lynne, 379, 411<br />

Keshishian, Haig, 9<br />

Kett, Lauren, 242<br />

Khan, Rehan M., 103, 310, 364, 428, 430, 432<br />

Khen, Miriam, 127<br />

Kicklighter, Cynthia, 269, 270<br />

Kida, Ikuhiro, 4<br />

Kidd, Judith, 304, 366<br />

Kidd, Kenneth K., 304, 366<br />

Kim, Joon, 11


Kim, K-O, 297<br />

Kim, Un-kyung, 191, 367, 433<br />

Kim, Yoon Tae, 320<br />

King, Michael S, 207<br />

Kings<strong>for</strong>d, Michael, 265<br />

Kinnamon, John C., 324, 325<br />

Kinnamon, Sue C., 83, 184, 326<br />

Kinsley, Elizabeth, 215<br />

Kleene, Nancy Koster, 282<br />

Kleene, Steven J., 282<br />

Klein, Jeffrey T., 69, 180<br />

Kleineidam, Christoph J., 218<br />

Klimbacher, Martina, 331<br />

Klock, Christopher, 119<br />

Klupp, Barbara, 175<br />

Knecht, Michael, 277<br />

Knipe, Cheryl, 341<br />

Kokrashvilli, Zaza, 189<br />

Kong, Ji Yong, 208<br />

Konnerth, Claus-Günther, 340<br />

Koo, Jae Hyung, 385<br />

Kornhuber, Johannes, 414<br />

Kosmidis, Efstratios, 7<br />

Koulakov, Alexei, 233<br />

Kratskin, Igor, 391<br />

Krautwurst, Dietmar, 362<br />

Kremser, Christian, 331<br />

Kroeze, Jan H. A., 299<br />

Krug, Patrick, 359<br />

Kubanek, Julia, 239<br />

Kurahashi, Takashi, 46<br />

Kusakabe, Yuko, 12<br />

Kuwasako, Takahiro, 29<br />

Kwon, Hyun J, 150<br />

Kwon, Hyung-Wook, 356<br />

Kwong, K., 346, 347<br />

Labra, Antoineta L, 34<br />

Laflam, Timothy, 119<br />

Lai, Peter C., 363<br />

Laing, David, 253<br />

Lancet, Doron, 127<br />

Landis, Basile N, 109, 339, 340<br />

Lane, Robert P, 36<br />

Langlois, Dominique, 312<br />

Lanier, Sarah, 19<br />

Larsson, Maria, 164<br />

Laska, Matthias, 249, 420<br />

Lawless, Harry T., 301<br />

le Coutre, Johannes, 330<br />

Lechner, Theresia, 331<br />

Lee, H-S, 297<br />

Lee, Jong-Ho, 320<br />

Lee, Virginia M.-Y., 379<br />

Lehmkuhle, Mark J, 6<br />

Lei, Hong, 230<br />

Leinders-Zufall, Trese, 150<br />

Leininger, Elizabeth Claire, 243<br />

119<br />

Leitch, Amanda, 56<br />

Lemon, Christian H, 93, 209<br />

Leon, Michael, 227, 228, 350<br />

Leonard, Ong, 126<br />

Lewcock, Joseph W., 132<br />

Li, Cheng Shu, 89<br />

Li, Cheng Xiang, 89<br />

Li, Cheng-shu, 278<br />

Li, Weiming, 157<br />

Li, Xia, 296<br />

Liggett, Rachel, 16<br />

Lilley, Sarah, 252<br />

Liman, Emily R, 190<br />

Lin, Dayu, 5<br />

Lin, Jeff, 163<br />

Lin, Weihong, 45, 98<br />

Linardopoulou, Elena, 352<br />

Linn, Charles, 217, 266<br />

Linschoten, Miriam R., 114<br />

Linster, Christiane, 60, 231<br />

Litaudon, Philippe, 158<br />

Liu, Dan, 190<br />

Liu, Guang, 48<br />

Liu, H., 347<br />

Liu, Hong-Xiang, 197<br />

Liu, Nian, 4, 229<br />

Liu, Weimin, 32, 33, 278<br />

Liu, Zhan, 295<br />

Longo, Melissa, 20<br />

Lorig, Tyler S., 100<br />

Loudon, Catherine, 262<br />

Lowe, Graeme, 223<br />

Lowry, Catherine A, 380<br />

LU, KUO-SHYAN S, 329<br />

Lu, Lu, 368<br />

Lucas, Nadia, 163<br />

Lucero, Mary T., 49, 254, 285<br />

Lui, S., 418<br />

Lundstrom, Johan N, 160, 277<br />

Lunkenheimer, Birgit, 414<br />

Lunkenheimer, Jens, 414<br />

Lyall, Vijay, 75, 76, 82, 188<br />

Lynch, David R, 335<br />

Ma, Jie, 223<br />

Macdonald, Jessica, 141<br />

Mackay-Sim, Alan, 336<br />

Macknin, Jonathan, 379<br />

Maggi, Leigh, 159<br />

Maher, Timothy J., 63<br />

Maier, Claudia, 331<br />

Mainland, Joel, 103, 310, 428, 430, 432<br />

Makino, Junshiro, 369<br />

Man, Orna, 127<br />

Manda, Saraswati, 385<br />

Mangold, Jamie Elizabeth, 322<br />

Mann, Wolf, 336<br />

Manzini, Ivan, 152


Margolis, Frank L., 150, 151, 385<br />

Margolis, Joyce W., 151<br />

Margolskee, Robert, 182, 183, 189, 199, 295, 370<br />

Marks, Lawrence E., 302<br />

Marlicz, Wojciech, 144<br />

Maroldt, Heike, 339<br />

Marshall, Katrina, 253<br />

Martin, Arthur, 235<br />

Martínez-Gómez, Margarita, 378<br />

Martinez-Marcos, Alino, 40<br />

Maruyama, Yutaka, 419<br />

Mascioli, Kirsten J., 214, 418<br />

Mast, Thomas, 400<br />

Matsumura, Kouichi, 240<br />

Matsunaga, Shigeki, 240<br />

Matsunami, Hiroaki, 128<br />

Matsunami, Momoka, 128<br />

Matsuoka, Masato, 41<br />

Mattes, Richard D., 30, 31, 410<br />

Maute, Christopher, 117, 167<br />

Max, Marianna, 182, 183, 295<br />

May, Darran, 376<br />

May, Olivia L., 321<br />

Maynard, Edwin M., 6<br />

Mbiene, Joseph-Pascal, 198, 203<br />

Mc Cune, Allicia, 112<br />

McBride, Kathleen, 248<br />

McBurney, Donald H., 161<br />

McCann, Jennifer, 61<br />

McCaughey, Stuart, 90<br />

McClintock, Martha, 244<br />

McClintock, Timothy S., 286, 287, 390<br />

McCluskey, Lynnette Phillips, 187<br />

McDermott, Ryan, 176<br />

McGlone, Francis, 65, 68<br />

McIntyre, Jeremy C., 286, 287<br />

McKee, Sherry, 305<br />

McLean, John Harvey, 61<br />

McNamara, Ann Marie, 60<br />

Mead, Kristina S., 261<br />

Mechaber, Wendy L., 267<br />

Mechtcheriakov, Sergei, 331<br />

Medler, Kathryn, 184<br />

Medrano, Juan F., 330<br />

Meisami, Esmail, 149, 387<br />

Mellinger, Craig, 303<br />

Mellon, DeForest, 405<br />

Menashe, Idan, 127<br />

Menco, Bert, 382<br />

Mennella, Julie A., 13, 14, 159, 214<br />

Meredith, M, 96, 275<br />

Meredith, Michael, 276, 392<br />

Mettenleiter, Thomas C., 175<br />

Meyerhof, Wolfgang, 191, 433<br />

Michel, William Craig, 147<br />

Michele, Gould, 117<br />

Miklavc, Pika, 54<br />

Miller, Ginger, 262<br />

120<br />

Miller, Lauren, 162<br />

Miller, Perry L., 229<br />

Miller-Sims, Vanessa, 406<br />

Mirich, Jennifer, 8<br />

Mistretta, Charlotte, 197, 204<br />

Mitzelfelt, Jeremiah D., 417<br />

Miura, Hirohito, 12<br />

Mizutani, Makoto, 369<br />

Moberg, Paul J., 122, 194, 332<br />

Mogyorosi, Andras, 178<br />

Mojet, Jos, 15, 22<br />

Møller, Per, 423<br />

Moncomble, Anne-Sophie, 312<br />

Moon, Cheil, 142<br />

Moon, Young Wha, 320<br />

Moore, Paul A., 52, 219, 234, 235, 236, 246, 258, 259, 260<br />

Moreland, Molly, 422<br />

Morrison, Edward E, 386<br />

Muhammed, Nizar, 333<br />

Munger, Steven D., 151, 213, 318<br />

Murphy, Claire, 251, 303, 309, 421, 422, 431<br />

Murray, Sarah G, 412<br />

Naidenko, Sergei, 271<br />

Naqvi, Ferheen, 167<br />

Neff, Jessica K., 341<br />

Nelson, Gina M, 323<br />

Nelson, Theodore M., 213, 318<br />

Neuhaus, Eva Maria, 130, 358<br />

Neves, Guilherme, 145<br />

Newcomb, Richard, 42<br />

Newman, Tera, 36, 352<br />

Newth, Angela, 319<br />

Nguyen, Eric A, 316<br />

Nicholson, Donald, 351<br />

Nickell, Tom, 288<br />

Nickles, Scott P., 238<br />

Nicolelis, Miguel A. L., 94<br />

Niessen, Heiner, 362<br />

Nighorn, Alan, 3, 44, 403<br />

Nikonov, Alexandre A., 97<br />

Ninomiya, Yuzo, 12, 73, 185, 189, 293, 327<br />

Nishiduka, Taichi, 29<br />

Nojima, Satoshi, 266<br />

Nomura, Mariko, 162<br />

Norita, Masao, 41<br />

Normann, Richard A, 6<br />

Northup, John K., 211<br />

Nosrat, Christopher A., 205<br />

Nosrat, Irina, 205<br />

O'Mahony, Michael, 297<br />

O'Malley, Stephanie S., 305<br />

Obermayer, Marieluise, 218<br />

Okabe, Masaru, 360<br />

Oland, Lynne A., 139, 255<br />

Olender, Tzviya, 127<br />

Ollinger, Fred, 225


Olsson, Mats J, 160, 277<br />

Olsson, Shannon Bryn, 245<br />

Opiekun, Richard, 116, 176, 338<br />

Ou, Guangshuo, 135<br />

Overton, Michael, 64<br />

Owens, Alicia, 422<br />

Pakstis, Andrew, 304, 366<br />

Pan, Yunfeng, 107<br />

Park, Daesik, 156<br />

Parker, Scott, 413<br />

Parsons, Almetra D., 64<br />

Pataramekin, Piyanuch Pahn, 387<br />

Pause, Betina M., 101<br />

Pawlik, Kate, 410<br />

Pazour, Gregory J., 134<br />

Pepino, Marta Yanina, 14, 214<br />

Perkins, Randa M., 64, 392<br />

Perreau-Linck, Elizabeth, 68<br />

Petrides, Michael, 102<br />

Phan, Tam-Hao T, 75, 76, 82, 188<br />

Philip, Shaji, 411<br />

Phillips, Nicola, 68<br />

Pirogovsky, Eva, 421, 422<br />

Pitovski, Dimitri Z., 307, 343<br />

Pittman, Dave Wayne, 28<br />

Pitts, Jason, 355<br />

Pixley, Sarah K., 282<br />

Plummer, Kim, 42<br />

Poor, Alexander, 100<br />

Porter, Jess, 430<br />

Porter, Jessica A., 103, 310, 428, 432<br />

Pouliot, Sandra, 424<br />

Prescott, John, 17<br />

Preston, Robin, 107<br />

Preti, George, 116<br />

Pribitkin, Edmund, 119, 342<br />

Puchalski, Diana, 56<br />

Puche, Adam C., 257, 394<br />

Qi, Ming, 361<br />

Quan, Wei, 40<br />

Quinn, Eric, 258<br />

Radil, Tomas, 66<br />

Raghow, Sandeep, 318<br />

Raguse, Jan-Dirk, 362<br />

Raitcheva, Denitza, 377<br />

Raman, Shanmugam Achi, 39<br />

Rasmussen, LEL, 42, 241, 242, 272, 314<br />

Raudenbush, Bryan, 111, 112, 113, 252<br />

Rawson, Ashley, 113<br />

Rawson, Nancy E., 291, 384<br />

Ray, Anandasankar, 131<br />

Ray, Koela, 106<br />

Reddy, Krishna, 411<br />

Reden, Jens, 177<br />

Redman, Christopher, 234<br />

121<br />

Reed, Danielle R., 191, 214, 296, 298, 304, 366, 418, 433<br />

Reed, Randall R., 132<br />

Reisenman, Carolina E, 138, 230<br />

Reisert, Johannes, 51, 279<br />

Reneerkens, Jeroen, 241<br />

Resasco, Marilina, 396<br />

Ressler, Kerry J, 59<br />

Restrepo, Diego, 2, 4, 45, 57, 98, 146, 292<br />

Ricardo, Torres, 421<br />

Richardson, Anne, 104<br />

Richter, Trevor A., 80, 81, 210<br />

Riddle, Scott W, 272<br />

Riffell, Jeffrey A., 130, 358, 359<br />

Rigdon, Megan, 100<br />

Rigsby, Christine' Spring, 187<br />

Rinberg, Dima, 225, 233<br />

Roalf, David R., 332<br />

Roberts, Cherie, 17<br />

Roberts, Richard, 128<br />

Robertson, Hugh M., 354<br />

Robinson, Alan M, 337, 348<br />

Robinson, Gene E., 372<br />

Rochlin, M. William, 201, 202<br />

Roeder, Hendrikje, 272<br />

Roelofs, Wendell, 266<br />

Roessler, Wolfgang, 218<br />

Rogers, Daniel H, 361<br />

Rogers, Jessica Lee, 404<br />

Rong, Minqing, 189<br />

Ronnett, Gabriele V., 142<br />

Roper, Stephen D., 80, 81, 210, 329, 419<br />

Rosen, David, 119, 342<br />

Rosenthal, Allison, 180<br />

Roskams, Angela Jane, 141, 351<br />

Ross, Susan R, 58<br />

Rossi, Ferdinando, 257<br />

Rossier, Olivier, 330<br />

Rothman, Douglas L., 4<br />

Rouby, Catherine, 427<br />

Rubenstein, Nicola, 430<br />

Rubrum, Adam M., 87<br />

Rugai, Nick, 91<br />

Rupp, Claudia I, 331<br />

Rützler, Michael, 43, 355<br />

Rybalsky, Konstantin, 118, 120, 121<br />

Sadamitsu, Chiharu, 73<br />

Saddoris, Michael P, 308<br />

Sainz, Eduardo, 211<br />

Saito, Harumi, 128<br />

Sakata, Yoko, 147<br />

Salcedo, Ernesto, 146<br />

Saldanha, Jason, 201<br />

Saleh, Maya, 351<br />

Sammeta, Neeraja, 287<br />

Samuelsen, Chad L., 276<br />

Sandra, Pouliot, 334


Sanematsu, Keisuke, 185<br />

Santos, Rosaysela, 11<br />

Sasaki, Hayato, 381<br />

Schaal, Benoist, 158, 312<br />

Schafer, Michele, 4<br />

Schannen, Andrew P., 281<br />

Scherer, Peter W, 342<br />

Schiffman, Susan S, 412<br />

Schild, Detlev, 152<br />

Schlabeck, Martin, 414<br />

Schlador, Michael, 352<br />

Schmidt, Katrina C, 324<br />

Schmidt, Manfred, 247, 389<br />

Schmiedeberg, Kristin, 362<br />

Schoenbaum, Geoffrey, 308<br />

Scholey, Jonathan M., 135<br />

Scholtz, Arne W, 331<br />

Scholz, Nathaniel, 376<br />

Schuckel, Julia, 375<br />

Schuler, Amanda, 113<br />

Schüler, Jenny, 164<br />

Schurian, Walter, 123<br />

Schwane, Katlen, 130, 358<br />

Schwarting, Gary, 377<br />

Schwob, James E, 143<br />

Scott, Alexander P., 157<br />

Scott, John W, 290, 402<br />

Seiden, Allen, 120<br />

Sethupathy, Praveen, 232<br />

Shabani, Shkelzen, 269<br />

Shah, Mussadiq, 333<br />

Shangari, Gopal Krishan, 429<br />

Shao, Zouyi, 394<br />

Shaw - Taylor, Ewurama E., 244<br />

Shepherd, Gordon M., 4, 229, 363, 395<br />

Sherrill, Lisa, 402<br />

Shetty, Ranjit S, 287<br />

Shigemura, Noriatsu, 73, 327<br />

Shigeoka, Yuko, 327<br />

Shingai, Tomio, 29<br />

Shipley, Michael T., 394, 399<br />

Shiraiwa, Takashi, 131<br />

Shirokova, Elena, 362<br />

Shoup, Melanie L., 161<br />

Shtoyerman, Eran, 224<br />

Shusterman, Dennis J., 169<br />

Sicard, Gilles, 158<br />

Silman, Eric, 11<br />

Silver, Wayne L., 171, 172, 173<br />

Simon, Sidney Arthur, 69, 94<br />

Simpson, P Jeanette, 142<br />

Singer, Michael, 363<br />

Sitthichai, Arie A., 49<br />

Siwicki, Kathleen K., 375<br />

Slack, Jay, 191, 433<br />

Slotnick, Burton, 45, 248<br />

Small, Dana M, 26, 68<br />

Smeets, Monique, 22, 116<br />

122<br />

Smith, Brian H., 62, 220, 262<br />

Smith, David V, 87, 89, 93, 186, 209<br />

Smith, Dean, 371<br />

Smith, James C., 294<br />

Smith, Jeffrey, 111, 112<br />

Smith, Melissa Anne, 287<br />

Smith, Michael B., 104<br />

Smith, Patrick, 294<br />

Snow, Joshua, 135<br />

Snyder, Derek, 304, 305, 306, 366<br />

Snyder, Lenore, 182, 295<br />

Sobel, Noam, 103, 310, 364, 428, 430, 432<br />

Soleimani, Manoocher, 282<br />

Sollars, Suzanne I, 206<br />

Sorensen, Peter W., 263, 264<br />

Spec, Andrej, 202<br />

Spector, Alan C., 77, 78, 79, 210, 319<br />

Speed, William, 304, 366<br />

Spehr, Jennifer, 168<br />

Spehr, Marc, 130, 168, 358<br />

Spielman, Andrew I., 330<br />

Spray, Kristina J, 181<br />

St. John, Steven J, 316, 368<br />

St. John, Steven James, 318<br />

Stakic, Josif, 401<br />

Stange, Gert, 267<br />

Stein, Heather, 138<br />

Steinbach, Molly Ann, 237<br />

Stengl, Monika, 375<br />

Stensmyr, Marcus C, 374<br />

Stepanyan, Ruben, 390<br />

Stepanyan, Tracy D., 390<br />

Stern, Shai, 127<br />

Stevens, David A., 301<br />

Stone, Leslie M, 326<br />

Strat<strong>for</strong>d, Jennifer M, 27<br />

Streeter, Sybil A., 161<br />

Stromberg, A. J., 347<br />

Stull, Kimberly, 421<br />

Suarez, Joseph C., 415<br />

Suchanek, Jessica, 401<br />

Sugita, Daigo, 293<br />

Suhr, Steven T., 221<br />

Sullivan, Daniel, 383<br />

Sullivan, Susan L., 211<br />

Swyers, Sarah, 386<br />

Szabo, Gabor, 394, 397<br />

Szeszko, Philip R, 331<br />

Tabert, Matthias H., 193<br />

Tai, P.C., 270, 388<br />

Takeuchi, Hiroko, 46<br />

Talamo, Barbara R., 48<br />

Tarun, Alice, 169<br />

Taylor, Gordon E., 350<br />

Taylor, Polly A, 294<br />

Taylor-Burds, Carol C., 416<br />

Tepper, Beverly J, 18


Tetsuya, Ookura, 12<br />

Tharp, Anilet A., 298<br />

Tharp, Christopher D., 23, 191, 298, 418, 433<br />

Thomas, Stacey, 325<br />

Thompson, Roger N., 64<br />

Thuerauf, Norbert, 414<br />

Tolbert, Leslie P, 137, 139<br />

Tordoff, Michael G, 296<br />

Touhara, Kazushige, 360<br />

Tourbier, Isabelle A., 335<br />

Trask, Barbara J., 36, 352<br />

Treloar, Helen B., 256<br />

Trojanowski, John, 379<br />

Trombley, Paul Q., 396<br />

Tucker, Kristal, 392<br />

Turetsky, Bruce I, 194, 332<br />

Turri, Linda, 245<br />

Ueda, Katsura, 200<br />

Uemura, Tadashi, 9<br />

Uflacker, Andre B, 207<br />

Umeh, Yvonne, 162<br />

Umezawa, Hidehiko, 369<br />

Urban, Lenka, 260<br />

Vaidehi, Nagarajan, 126<br />

Vainius, Aldona, 66<br />

Vaishnav, R. A., 346, 347<br />

Valentincic, Tine, 54, 153<br />

Valentine, Megan, 105<br />

van der Goes van Naters, Wynand, 131<br />

Van Houten, Judith, 105, 106, 107, 124, 283<br />

Veldhuizen, Maria Geraldine, 299<br />

Verret, Travis, 280<br />

Vetter, Monica L., 10<br />

Vickers, Neil J., 55, 217<br />

Viera, Erin, 421<br />

Vieyra, Michelle, 353<br />

Vigers, Alison, 83<br />

Vilbig, Ryan, 201<br />

Villarreal, Rebecca, 162<br />

Vinnikova, Anna K, 76, 82, 188<br />

Visegrady, Andras, 393<br />

Vogalis, Fivos, 254, 285<br />

Vogt, Richard G., 136, 268, 353<br />

Vosshall, Leslie B, 250, 357<br />

Voznessenskaya, Vera, 271<br />

Vrieze, Lance A., 263<br />

Vucinic, Dejan, 7<br />

Wada, Yumiko, 369<br />

Wakabayashi, Yoshihiro, 37<br />

Wakisaka, Sayoshi, 200<br />

Walch, Thomas, 331<br />

Walker, Dianne, 415<br />

Walker, James, 415<br />

Walker, Megan, 352<br />

Walters, Eric, 198<br />

123<br />

Walworth, Ellen, 140<br />

Wang, Dalton, 32, 278<br />

Wang, Hong, 183<br />

Wang, Jian-li, 104<br />

Wang, Jing, 72<br />

Warren, Craig B., 361<br />

Watanabe, Akira, 199<br />

Waters, Robert S, 89<br />

Watson, Erika L, 412<br />

Weenen, Hugo, 22, 409<br />

Weeraratne, Shyamal Dilhan, 105, 124<br />

Weiler, Elke, 345<br />

Weinrich, Karl, 63<br />

Welge-Lüssen, Antje, 165<br />

Westberry, Jenne, 96, 276<br />

Westburg, Anna M., 416, 417<br />

Westerterp, Margriet, 22<br />

Wetzel, Christian, 168, 175<br />

White, Joel E., 125, 289, 365<br />

White, Theresa L., 20<br />

Whitt, Robert, 415<br />

Wifall, Timothy C., 416<br />

Wilkes, Annie, 422<br />

Willander, Johan Magnus, 164<br />

Willecke, Klaus, 362<br />

Williams, Eleanor, 352<br />

Williams, Jayme Melissa, 404<br />

Williams, Lloyd B., 125, 365<br />

Williams, Robert W., 368<br />

Wilson, Caroline, 403<br />

Wilson, Donald, 226<br />

Wilson, Donald A., 95, 99<br />

Wilson, Patricia, 425, 426<br />

Wilson, Tracy, 110<br />

Winnig, Marcel, 191<br />

Wirkus, Eric, 154<br />

Wirsig-Wiechmann, Celeste Renee, 155<br />

Wise, Paul, 66<br />

Witman, George B., 133<br />

Wolf, Mary, 219, 258<br />

Wolfensberger, Markus, 165<br />

Wong, Allan M, 72<br />

Woo, Cynthia C., 350<br />

Woo, Jinkyeung, 143<br />

Wooding, Stephen, 367<br />

Wright, Geraldine, 62<br />

Wright, James C, 386<br />

Wulff, Christian, 423<br />

Wysocki, Charles, 66<br />

Wysocki, Linda M., 291<br />

Xu, Fuqiang, 4, 229<br />

Xu, PingXi, 371<br />

Yamada, Ayako, 293<br />

Yamashita, Shizuya, 29<br />

Yamazaki, Kunio, 58<br />

Yamout, Adam, 202


Yang, Hsiuchin, 270<br />

Yang, Qing X., 104<br />

Yang, Ruibiao, 324, 325<br />

Yano, J., 105, 107<br />

Yano, Junji, 106, 124<br />

Yao, Ying, 9<br />

Yasumatsu, Keiko, 73, 189, 327<br />

Yau, King-Wai, 50, 51, 279<br />

Ye, Mi Kyung, 87<br />

Yee, Karen K., 384<br />

Yomogida, Kentaro, 360<br />

Yoshida, Ryusuke, 73, 185<br />

Young, Janet, 36<br />

Young, Janet M., 352<br />

Yu, Katie, 157<br />

Yu, Tun-Tzu, 286, 287<br />

Yue, Esther, 60<br />

Yun, Sang Seon, 157<br />

Yurchenko, Olga, 155<br />

Zald, David H., 68<br />

Zatorre, Robert, 101<br />

Zatorre, Robert J, 102<br />

Zelano, Christina, 103, 310, 428, 430, 432<br />

Zellner, Debra, 413<br />

Zhang, Chunbo, 2, 292<br />

Zhang, Jianhua, 282<br />

Zhao, Haiqing, 151<br />

Zhao, Kai, 342<br />

Zhao, Liqiang, 18<br />

Zhu, Mingyan, 398<br />

Zhukovskaya, Marianna, 107<br />

Zielinski, Barbara S., 157, 284<br />

Zimmer, Richard K., 130, 358, 359<br />

Zoladz, Phillip, 252<br />

Zou, Zhihua, 311<br />

Zucker, Jacob, 145<br />

Zufall, Frank, 150<br />

Zuker, Charles S., 71<br />

Zulandt, Thomas John, 258<br />

Zuri, Ido, 38, 273<br />

Zwiebel, Laurence J., 43, 355, 356<br />

124

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!