08.06.2015 Views

Abstracts - Association for Chemoreception Sciences

Abstracts - Association for Chemoreception Sciences

Abstracts - 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.

AChemS<br />

<strong>Association</strong> <strong>for</strong> <strong>Chemoreception</strong> <strong>Sciences</strong><br />

ABSTRACTS<br />

AChemS XXXV<br />

The <strong>Association</strong> <strong>for</strong> <strong>Chemoreception</strong> <strong>Sciences</strong><br />

35th Annual Meeting<br />

April 17-20, 2013<br />

Hyatt Regency<br />

Huntington Beach, CA<br />

1


PepsiCo is proud to support<br />

AChemS<br />

and its commitment to advancing an understanding of the chemical senses<br />

At PepsiCo, Per<strong>for</strong>mance with Purpose means delivering sustainable growth by investing<br />

in a healthier future <strong>for</strong> people and our planet. We will continue to build a portfolio of<br />

enjoyable and wholesome foods and beverages, find innovative ways to reduce the use<br />

of energy, water and packaging, and provide a great workplace <strong>for</strong> our associates.<br />

Because a healthier future <strong>for</strong> all people and our planet<br />

means a more successful future <strong>for</strong> PepsiCo.<br />

www.pepsico.com


Table of Contents<br />

2013 AChemS Meeting Sponsors & Corporate Members. .............................. 5<br />

2013 AChemS Meeting Exhibitors. .............................................. 6<br />

2013 Awardees ........................................................... 7<br />

Committees .............................................................. 9<br />

Oral <strong>Abstracts</strong> ........................................................... 10<br />

Posters. ............................................................... 30<br />

Author Index ........................................................... 130<br />

Program at a Glance (Visual) ................................................ 140<br />

Please Note: Filming and photographing presentations (plat<strong>for</strong>m and poster) is prohibited unless the<br />

presenting author has granted permission.<br />

4


AChemS<br />

<strong>Association</strong> <strong>for</strong> <strong>Chemoreception</strong> <strong>Sciences</strong><br />

extends special thanks and appreciation <strong>for</strong> grant support from:<br />

The National Institute on Deafness and<br />

Other Communication Disorders<br />

and the<br />

National Institute on Aging, NIH<br />

The <strong>Association</strong> <strong>for</strong> <strong>Chemoreception</strong> <strong>Sciences</strong> is also grateful<br />

<strong>for</strong> the generous support of its corporate sponsors:<br />

Diamond Sponsors<br />

Platinum Sponsor<br />

Silver Sponsors<br />

Other Sponsors<br />

A special thank you to Ghislaine Polak and the late Ernest Polak <strong>for</strong><br />

supporting the Polak Young Investigators Awards and the<br />

Junior Scientist Travel Awards.<br />

The <strong>Association</strong> <strong>for</strong> <strong>Chemoreception</strong> <strong>Sciences</strong> thanks our<br />

Corporate Members <strong>for</strong> their support.<br />

5


2013 Annual Meeting Exhibitors<br />

EXHIBIT HOURS<br />

Thursday, April 18<br />

8:00 am – 12:00 pm<br />

Friday, April 19<br />

8:00 am – 12:00 pm<br />

Saturday, April 20<br />

8:00 am – 12:00 pm<br />

Ox<strong>for</strong>d University Press<br />

Ox<strong>for</strong>d University Press is a publisher of some<br />

of most prestigious books and journals in the world. They include Chemical<br />

Senses, the official journal of ACHEMS, ECRO and JASTS. Visit our stand<br />

to pick up gratis copies of our journals, or go to www.ox<strong>for</strong>djournals.org<br />

to read a free issue online.<br />

Company representative: Jennifer Boyd<br />

Sensonics, Inc.<br />

Sensonics International provides the medical, scientific and industrial<br />

communities with the best smell and taste tests <strong>for</strong> assessing<br />

chemosensory function.<br />

Company representative: Kyra Milnamow<br />

6


2013 Awardees<br />

35 th Annual Givaudan Lectureship<br />

Bill Hansson, PhD, Max Planck Institute <strong>for</strong> Chemical Ecology, Jena, Germany<br />

19 th Annual Ajinomoto Award to Promising Young Researcher in the Field of Gustation<br />

Peihua Jiang, PhD, Monell Chemical Senses Center, Philadelphia, PA, USA<br />

International Flavors and Fragrances Award <strong>for</strong> Outstanding Research on the Molecular Basis of Taste<br />

Wolfgang Meyerhof, PhD, German Institute of Human Nutrition, Nuthetal, Germany<br />

22 nd Annual Moskowitz Jacobs Award <strong>for</strong> Research in Psychophysics of Human Taste and Smell<br />

Wen Li, PhD, University of Wisconsin-Madison, Madison, WI, USA<br />

Max Mozell Award <strong>for</strong> Outstanding Achievement in the Chemical Senses<br />

Stuart Firestein, PhD, Columbia University, New York, NY, USA<br />

The AChemS Young Investigator Award <strong>for</strong> Research in Olfaction<br />

Haiqing Zhao, PhD, Johns Hopkins University, Baltimore, MD, USA<br />

The Don Tucker Memorial Award (2012 Awardee)<br />

Cecil “Jake” Saunders, The University of Colorado, Aurora, CO, USA<br />

The Polak awards are funded by the Elsje Werner-Polak Memorial Fund in memory of our niece<br />

gassed by the Nazis in 1944 at age 7:<br />

Ghislaine Polak and the late Ernest Polak<br />

Polak Young Investigator Award Recipients<br />

Pablo Chamero, University of Saarland, Homburg, Germany<br />

Adam Dewan, Northwestern University, Evanston, IL, USA<br />

Dany Gaillard, University of Colorado Denver, Denver, CO, USA<br />

Ahmad Jezzini, SUNY, Stony Brook, NY, USA<br />

Kathrin Ohla, German Institute of Human Nutrition, Nuthetal, Germany<br />

Markus Rothermel, University of Utah School of Medicine, Salt Lake City, UT, USA<br />

7


2013 Awardees<br />

We are pleased to announce that five 2013 Polak Junior Scientist<br />

Travel Awards were given <strong>for</strong> this year’s Meeting.<br />

AChemS Travel Fellowships <strong>for</strong> Diversity Recipients:<br />

Funded by a generous grant from the National Institute on Deafness and<br />

Other Communication Disorders and the National Institute on Aging, NIH<br />

Norma Castro, San Diego State University, San Diego, CA, USA<br />

Deandrea Ellis, Columbia University, New York, NY, USA<br />

Melissa Haley, Stony Brook University, Stony Brook, NY, USA<br />

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

Yasmin Marrero, Brandeis University, Waltham, MA, USA<br />

Jacquelyn Szajer, San Diego State University, San Diego, CA, USA<br />

Cedric Uytingco, University of Maryland School of Medicine,<br />

Baltimore, MD, USA<br />

AChemS Student Housing and Travel Award Recipients<br />

Funded by the Polak Foundation: Ghislaine Polak and the late Ernest Polak<br />

Tobias Ackels<br />

Sarah Ashby<br />

Dylan Barnes<br />

Muhammad Binyameen<br />

Anna Blumrich<br />

Yvonne Brünner<br />

Nadia Byrnes<br />

David Castillo<br />

Jennifer Chen<br />

Kepu Chen<br />

Adam Clark<br />

David Collins<br />

Rachel Dana<br />

Olga Escanilla<br />

Elena Flohr<br />

Tomomi Fujimaru<br />

Jessica Gaby<br />

Monika Gorin<br />

Marco Guarneros<br />

Claudia Haering<br />

Shigeki Inoue<br />

Mavis Irwin<br />

Joseph Jones<br />

Marley Kass<br />

Kurt Krosnowski<br />

Devaki Kumarhia<br />

Trina Lapis<br />

Lucas Lemasters<br />

Rosemary Lewis<br />

Yan Liu<br />

SiWei Luo<br />

Ali Magableh<br />

Shahid Majeed<br />

Amanda Maliphol<br />

Yasmin Marrero<br />

Amanda McKenna<br />

Dulce Minaya<br />

Andrew Moberly<br />

Sierra Moore<br />

Melissa Murphy<br />

Stephanie Oleson<br />

In Jun Park<br />

Shristi Rawal<br />

Sébastien Romagny<br />

Kimberly Smith<br />

Xue Sun<br />

Jacquelyn Szajer<br />

Anna K Talaga<br />

William Tewalt<br />

Darcy Trimpe<br />

Cedric Uytingco<br />

Megha Vasavada<br />

Crystal Wall<br />

Pei Xu<br />

Wendy Yoder<br />

Shaohua Zhao<br />

Harriët Zoon<br />

Logo Award Winner<br />

Cedric Uytingco, University of Maryland, School of Medicine, Baltimore, MD, USA<br />

8


Committees<br />

ACHEMS EXECUTIVE COMMITTEE 2012-2013<br />

President Alan Spector, PhD Florida State University<br />

Past President Timothy McClintock, PhD University of Kentucky<br />

Senior Advisor Don Wilson, PhD Nathan Kline Institute and<br />

NYU School of Medicine<br />

President Elect John Glendinning, PhD Barnard College,<br />

Columbia University<br />

Secretary Julie Mennella, PhD Monell Chemical<br />

Senses Center<br />

Membership Chair Dana Small, PhD JB Pierce Laboratory/<br />

Yale University<br />

Program Chair Paul Breslin, PhD Rutgers University and<br />

Monell Chemical<br />

Senses Center<br />

Treasurer Joe Travers, PhD The Ohio State University<br />

Program Chair Elect Steven Munger, PhD University of Maryland<br />

School of Medicine<br />

Sr. Councilor Haiqing Zhao, PhD Johns Hopkins<br />

University<br />

Jr. Councilor Christiane Linster, PhD Cornell University<br />

ACHEMS PROGRAM COMMITTEE 2012-2013<br />

Paul Breslin (Chair)<br />

Robert Anholt<br />

Maik Behrens<br />

Nirupa Chaudhary<br />

Denise Chen<br />

Jean-Francois Cloutier<br />

Diego Rodriguez Gil<br />

Don Katz<br />

Johan Lundstrom<br />

Minghong Ma<br />

Bettina Malnic<br />

John McGann<br />

Steve Munger<br />

Venky Murthy<br />

Brian Smith<br />

Mark Stopfer<br />

Lisa Stowers<br />

Ali Ventura<br />

Paul Wise<br />

MEETING EVALUATION<br />

The meeting evaluation is available online this year. Please visit www.achems.org to give us your<br />

feedback on the meeting. Your input helps AChemS’ leadership continue to offer quality annual meetings<br />

and member services.<br />

9


Oral <strong>Abstracts</strong><br />

#1 GIVAUDAN LECTURE:<br />

NON-MODEL MODELS IN OLFACTION<br />

#2 SYMPOSIUM: THE STRUCTURAL<br />

BASIS OF CHEMOSENSORY SIGNALING<br />

Bill S. Hansson<br />

Max Planck Institute <strong>for</strong> Chemical Ecology, Jena, Germany<br />

Our knowledge regarding olfactory structure and function has<br />

taken a quantum leap since the characterization of putative<br />

olfactory receptors both in vertebrates and insects. Still, the<br />

understanding of both ecological and evolutionary processes is<br />

lagging behind. What do different odors really mean to animals,<br />

and how has the system evolved to allow them to respond<br />

behaviorally in a relevant manner. To understand these topics<br />

better we use both a model insect (Drosophila melanogaster),<br />

and a number of non-model arthropods (moths, ants, primitive<br />

insects, land-living crustaceans).<br />

I will spend a short introductory part on recent progress in<br />

our work on drosophilid flies, touching on investigations of<br />

hard-wired smell-driven behavior. The rest of my talk I will<br />

devote to non-model arthropods.<br />

In the desert ant, Cataglyphis <strong>for</strong>tis, we have built on the classic<br />

investigations by the group of Rüdiger Wehner in Zürich. They<br />

demonstrated the amazing ability of these tiny insects to find<br />

their way home in a salt desert using vector-based orientation.<br />

In our investigations we show how the ants use odor cues to<br />

pinpoint their nest entrance, but how this orientation is, in a<br />

life-saving fashion, weighed against vector orientation. We also<br />

show how the ants use olfactory cues to find their main source<br />

of food; dead insects, in a highly efficient way, and how this<br />

smell-based food search is independent of the home vector.<br />

To investigate the evolution of arthropod olfactory systems<br />

we make use of both highly primitive, archeopteran insects<br />

(in comparison to their neopteran relatives), and of land-living<br />

crustaceans that have entered the terrestrial environment<br />

during the last five million years. Using a combination of<br />

transcriptomics, electrophysiology, morphology and bioassays<br />

we show that the neopteran divide coincides with a drastic<br />

development in the insect olfactory system, and that some<br />

land-living crustaceans have developed enormous central<br />

olfactory systems, while others seem to have abandoned<br />

olfaction.<br />

G protein coupling GPCRs: structural and functional<br />

insights into reciprocal G protein and GPCR interactions<br />

Roger K. Sunahara<br />

University of Michigan Medical School Ann Arbor, MI, USA<br />

Recent advances in the structural biology of G protein-coupled<br />

receptors have helped to unravel the intricacies of ligand binding.<br />

Similarly structural and biochemical analyses of heterotrimeric<br />

G proteins have affirmed our understanding of the mechanism<br />

underlying effector interactions and GTPase activity. The recent<br />

crystal structure of a prototypic GPCR, the beta 2<br />

-adrenergic<br />

receptor (beta 2<br />

AR), in a complex with the stimulatory G protein,<br />

Gs, trapped in its nucleotide-free state, has now provided models<br />

<strong>for</strong> activation of G proteins by GPCRs. These data have helped<br />

to delineate how hormone binding to GPCRs leads to GDP<br />

release on G proteins, the principle step that precedes GTP<br />

binding and G protein activation. The crystal structure, together<br />

with data from single particle reconstructions by electron<br />

microscopy and deuterium exchange mass spectrometry, reveal<br />

dramatic changes in the G protein alpha-subunit. The structural<br />

data also suggest that G proteins may allosterically regulate the<br />

receptor by stabilizing a closed con<strong>for</strong>mation on the extracellular<br />

face of the receptor. Radioligand binding analyses suggest<br />

that G protein coupling slows ligand dissociation, consistent<br />

with the observed structural changes in the extracellular face.<br />

Such structural changes account <strong>for</strong> the slower observed ligand<br />

dissociation rates and likely account <strong>for</strong> G protein-dependent<br />

high affinity agonist binding. Together these data support a<br />

plausible model <strong>for</strong> the mechanism <strong>for</strong> receptor-mediated<br />

nucleotide exchange, G protein activation and agonist binding.<br />

Acknowledgements: GM083118<br />

#3 SYMPOSIUM: THE STRUCTURAL<br />

BASIS OF CHEMOSENSORY SIGNALING<br />

Structural Determinants of TRPV Channel Activation<br />

and Desensitization<br />

Rachelle Gaudet<br />

Harvard University Cambridge, MA, USA<br />

ORAL ABSTRACTS<br />

My lab is broadly interested in the mechanisms of signaling<br />

and transport across cellular membranes. Much of our research<br />

centers on TRP channels and their role in sensory perception.<br />

We focus on temperature-sensitive ion channels, particularly<br />

TRPV1 and TRPA1. TRP channels are challenging structural<br />

biology targets because they are large multidomain eukaryotic<br />

membrane proteins and are not naturally abundant. We<br />

take complementary approaches to obtain structural and<br />

functional in<strong>for</strong>mation on TRP channels. One strategy is to<br />

divide and conquer: determine crystal structures of isolated<br />

domains of TRP channels. The results can then combined with<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

10


genetic, biochemical and physiological data to advance our<br />

understanding of TRP channel function. It also provides valuable<br />

insights as we tackle the structure determination of whole TRP<br />

channels. Acknowledgements: NIH grant R01GM081340 to RG.<br />

#4 SYMPOSIUM: THE STRUCTURAL<br />

BASIS OF CHEMOSENSORY SIGNALING<br />

Atomistic Structures of human olfactory and taste receptors<br />

William A Goddard, Soo-Kyung Kim, Ravinder Abrol<br />

Materials and Process Simulation Center, Cali<strong>for</strong>nia Instute of<br />

Technology Pasadena, CA, USA<br />

Olfactory receptors (ORs) responsible <strong>for</strong> mediating the sense<br />

of smell are G protein-coupled receptors (GPCRs). Through<br />

differential interactions the ~400 ORs allow us to recognize<br />

countless ligands. However little is known about the specific<br />

interactions responsible. We use first principles (non homology)<br />

methods to predict the ensemble of 24 low energy structures <strong>for</strong><br />

the human OR1G1 (hOR1G1) using the GEnSeMBLE (GPCR<br />

Ensemble of Structures in Membrane BiLayer Environment)<br />

method. The final best structure has strong interhelical<br />

H-bonding networks in TMs 1-2-7 (N1.50, D2.50 and N7.49),<br />

and TMs 2-4 (N2.45 and W4.50), which have been observed<br />

in x-ray and predicted structures <strong>for</strong> some members of the<br />

class A (rhodopsin family) GPCRs. We also find a salt-bridge<br />

between the conserved D3.49 and K6.30 in the D(E)RY region,<br />

associated with the inactive <strong>for</strong>m <strong>for</strong> class A GPCRs. In addition,<br />

a hOR specific interaction was <strong>for</strong>med between the conserved<br />

D/E3.39 and H6.40 in hORs, which might be an important<br />

con<strong>for</strong>mational constraint in all hORs since they are highly<br />

conserved ( ~94 and 98% respectively) among all hORs. We<br />

expect our predicted 3D structures <strong>for</strong> hOR1G1 to provide a<br />

guide in understanding the structural patterns <strong>for</strong> all hORs.<br />

This might help design better odorants <strong>for</strong> foods and perfumes<br />

and improved drugs <strong>for</strong> neurodegenerative disorders, such as<br />

anosmia (inability to detect odors) or hyposmia (decreased ability<br />

to detect odors). The structure and conserved elements of the<br />

hORs are compared to our recent prediction of the atomistic<br />

structure <strong>for</strong> the human TAS2R38 Bitter Receptor.<br />

TAS2R16 mutation library with defined point mutations at every<br />

amino acid in the receptor and screened each mutant’s functional<br />

activity in response to unique ligands, both agonists and<br />

inhibitors, using a Ca 2+ -flux signaling assay. We identified amino<br />

acids that, when substituted, abrogate signaling <strong>for</strong> all agonists,<br />

as well as residues important <strong>for</strong> the activity of only specific<br />

ligands. These critical residues define both the binding sites <strong>for</strong><br />

various TAS2R16 ligands as well as motifs important <strong>for</strong> signal<br />

transduction, and will help guide the structural understanding<br />

of bitter taste receptor function. Acknowledgements: NIH<br />

DC002995 NIH DC010105 NIH GM072379<br />

#7 INDUSTRY SYMPOSIUM:<br />

TASTE AND SMELL IN TRANSLATION:<br />

APPLICATIONS FROM BASIC RESEARCH<br />

Individual differences in oral fat detection and their<br />

health implications<br />

Richard Mattes<br />

Purdue University, West Lafayette IN 47907 USA<br />

The fat content of foods and the oral perception of this fat<br />

influence food choice and a wide range of physiological<br />

responses with health implications. Triacylglycerol, the<br />

predominant <strong>for</strong>m of dietary fat, generally contributes positive<br />

somatosensory attributes to foods and are viewed as orexigenic.<br />

In contrast, non-esterified fatty acids, likely taste stimuli,<br />

typically elicit negative hedonic responses and are anorexigenic.<br />

The latter also exerts unique effects on cephalic, intestinal<br />

and post-absorptive lipid metabolism. However, individual<br />

variability in taste and physiological responses is extremely high<br />

raising question about distinct groups of fat tasters/responder<br />

This presentation will examine possible sources of individual<br />

variability such as learning, BMI status, age, gender and stimuli<br />

properties to facilitate better study designs and interpretation of<br />

the functions of “fat taste.”<br />

#8 INDUSTRY SYMPOSIUM:<br />

TASTE AND SMELL IN TRANSLATION:<br />

APPLICATIONS FROM BASIC RESEARCH<br />

ORAL ABSTRACTS<br />

#5 SYMPOSIUM: THE STRUCTURAL<br />

BASIS OF CHEMOSENSORY SIGNALING<br />

Comprehensive mapping of functional sites <strong>for</strong> agonists and<br />

inhibitors of the bitter taste receptor TAS2R16<br />

Joseph B. Rucker<br />

Integral Molecular Philadelphia, PA, USA<br />

Bitter tastes are detected at the cellular level by a diverse family<br />

of taste receptors (TAS2Rs) belonging to the G protein-coupled<br />

receptor (GPCR) superfamily. The lack of direct structural<br />

in<strong>for</strong>mation <strong>for</strong> TAS2Rs (and most GPCRs) limits our<br />

understanding of their structural features responsible <strong>for</strong> ligand<br />

binding and signaling. Using a high-throughput approach called<br />

Shotgun Mutagenesis Mapping, we created a comprehensive<br />

Mechanisms <strong>for</strong> Fat Taste<br />

Timothy Gilbertson<br />

Department of Biology, Utah State University, Logan UT 84322 USA<br />

It has been well over a decade since our laboratory first identified<br />

the ability of free fatty acids to activate mammalian taste<br />

receptors cells, consistent with there being a “taste of fat”. Since<br />

that time, the ability of fatty acids to act as the proximate stimuli<br />

<strong>for</strong> fat taste has been validated in a number of species spanning<br />

the molecular, cellular and behavioral levels. We have recently<br />

identified several novel fatty acid-activated G protein-coupled<br />

receptors that, in conjunction with the fatty acid binding protein<br />

CD36, allow the recognition of chemically distinct classes of<br />

free fatty acids. This presentation will summarize what is known<br />

about the receptors and transduction pathway <strong>for</strong> free fatty acids.<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

11


#9 INDUSTRY SYMPOSIUM:<br />

TASTE AND SMELL IN TRANSLATION:<br />

APPLICATIONS FROM BASIC RESEARCH<br />

Insights from olfactory receptor screening<br />

Joel Mainland<br />

Monell Chemical Senses Center, Philadelphia PA 19104 USA<br />

A major advances in the taste field in recent decades was the<br />

identification of receptors that mediate taste. Expression of these<br />

receptors in heterologous cell-based assays has allowed scientists<br />

in both academia and industry to screen <strong>for</strong> novel taste agonists,<br />

antagonists, and modifiers. Similar studies in olfactory receptors<br />

have lagged behind the taste field due to the size of the receptor<br />

family as well as difficulties in expressing the receptors in<br />

heterologous cells. In this talk we will explore the current state of<br />

the science in olfactory receptor screening, relationships between<br />

odorant structure and odor quality, and the identification of<br />

agonists, antagonists and modifiers <strong>for</strong> olfactory receptors.<br />

#10 INDUSTRY SYMPOSIUM:<br />

TASTE AND SMELL IN TRANSLATION:<br />

APPLICATIONS FROM BASIC RESEARCH<br />

Mechanisms of olfactory adaptation<br />

Haiqing Zhao<br />

Johns Hopkins University, Baltimore MD 21218 USA<br />

Olfactory receptor cells exhibit reduced sensitivity upon<br />

prolonged or repeated odor exposure––a phenomenon known as<br />

adaptation. Adaptation at the cellular level is thought to underlie,<br />

at least in part, the perceptual desensitization of an odor<br />

over time. In vertebrates, several calcium-dependent feedback<br />

mechanisms have been proposed to account <strong>for</strong> adaption of<br />

olfactory receptor cells. Recent studies using molecular genetic<br />

approaches that allow selective disruption of these calciumdependent<br />

mechanisms have provide new insight into how<br />

olfactory adaptation may occur.<br />

#12 PRESIDENTIAL SYMPOSIUM:<br />

GUT PEPTIDE INTERACTIONS BETWEEN<br />

TASTE, FEEDING, AND REWARD<br />

Bariatric surgery and appetite<br />

Carel Le Roux<br />

Experimental Pathology, UCD Conway Institute, School of Medicine<br />

and Medical Science, University College Dublin, Ireland<br />

A good model to investigate appetite reduction in humans and<br />

rodents with associated major weight loss is bariatric surgery.<br />

Gastric bypass, but not gastric banding caused increased<br />

postprandial PYY and GLP-1 favouring enhanced satiety.<br />

An early and exaggerated insulin response mediates improved<br />

glycaemic control. The rodent model of bypass showed elevated<br />

PYY, GLP-1 and gut hypertrophy compared with sham-operated<br />

rats. Moreover, exogenous PYY reduced food intake while<br />

blockade of endogenous PYY increased food intake.<br />

A prospective follow-up human study of gastric bypass showed<br />

progressively increasing PYY, enteroglucagon, and GLP-1<br />

responses associated with enhanced satiety. Blocking these<br />

responses in animal and human models leads to increased food<br />

intake. Thus, following gastric bypass, a pleiotrophic endocrine<br />

response may contribute to improved glycaemic control,<br />

appetite reduction, and long-term lowering of body weight.<br />

We have shown that changes occur in the sensory, reward<br />

and physiological domains of taste that may mechanistically<br />

contribute to the alterations in food preferences after gastric<br />

bypass. The sustained nature of weight loss, reduced appetite<br />

and shifts in food preferences may be explained by gut adaptation<br />

and chronic hormone elevation.<br />

#13 PRESIDENTIAL SYMPOSIUM:<br />

GUT PEPTIDE INTERACTIONS BETWEEN<br />

TASTE, FEEDING, AND REWARD<br />

Common Mechanisms of Alimentary Chemosensation:<br />

Implications <strong>for</strong> Taste, Ingestion and Glucose Homeostasis<br />

Steven D. Munger 1,2<br />

1<br />

University of Maryland School of Medicine, Department of Anatomy<br />

and Neurobiology Baltimore, MD, USA, 2 University of Maryland<br />

School of Medicine, Department of Medicine Baltimore, MD, USA<br />

The last two decades has seen a growing recognition that a<br />

common molecular toolkit is employed along the length of the<br />

alimentary canal to detect and respond to nutrients. For example,<br />

many of the same proteins that are critical <strong>for</strong> recognizing sweet,<br />

bitter and umami taste stimuli in the mouth, including T1R and<br />

T2R taste receptors and the transduction proteins a-gustducin<br />

and TRPM5, are found throughout the gastrointestinal (GI) tract<br />

and associated organs. Similarly, taste buds express a number of<br />

neuropeptides that are perhaps best understood in other systems<br />

as endocrine factors that impact nutrient metabolism and/or<br />

ingestive behaviors. A better understanding of the molecular<br />

mechanisms that couple nutrient detection to peptide secretion in<br />

gustatory and GI tissues could lead to the identification of new<br />

pharmacological targets <strong>for</strong> impacting ingestion, satiety, nutrient<br />

assimilation or glycemic control. I will discuss our recent studies<br />

in rodents, including animals deficient in specific taste receptor<br />

subunits or receiving Roux-en-Y gastric bypass, that provide<br />

new insights into the mechanisms of nutrient response in the<br />

mouth and gut and the role of these mechanisms in taste coding,<br />

post-ingestive nutrient response, and the regulation of glucose<br />

homeostasis. Acknowledgements: NIDCD (DC010110), Tate &<br />

Lyle Americas, Ajinomoto Amino Acid Research Program<br />

ORAL ABSTRACTS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

12


#14 PRESIDENTIAL SYMPOSIUM:<br />

GUT PEPTIDE INTERACTIONS BETWEEN<br />

TASTE, FEEDING, AND REWARD<br />

GLP-1 neurons: A link between gastrointestinal satiation<br />

signaling and food reward?<br />

Diana L. Williams<br />

Psychology Department & Program in Neuroscience, Florida State<br />

University Tallahassee, FL, USA<br />

Central glucagon-like peptide 1 receptor (GLP-1R) stimulation<br />

suppresses food intake, and hindbrain GLP-1 neurons project<br />

to numerous feeding-relevant brain regions. One such region is<br />

the nucleus accumbens (NAc), which plays a role in reward and<br />

motivated behavior. Using immunohistochemical and retrograde<br />

tracing techniques in rats, we identified a robust projection from<br />

GLP-1 neurons in the nucleus of the solitary tract to the NAc<br />

core. Our pharmacologic studies then provided evidence that<br />

exogenous and endogenous GLP-1R stimulation in the NAc<br />

core reduces food intake. The NAc is known to influence food<br />

palatability and motivation, while GLP-1 neurons have been<br />

implicated in mediation of meal-related gastrointestinal signals.<br />

We hypothesize that GLP-1 released in NAc in response to<br />

nutrient ingestion contributes to meal-induced satiation, and that<br />

GLP-1 action in NAc reduces the hedonic value of food.<br />

To understand the role of endogenously released GLP-1 in the<br />

NAc core, we have been per<strong>for</strong>ming detailed behavioral analysis<br />

of the effects of blockade of GLP-1R at this site. We find that<br />

injection of the GLP-1R antagonist exendin 9-39 (Ex9) into the<br />

NAc core selectively increases meal size and blunts the satiating<br />

effects of nutrients. In rats consuming sucrose solutions, NAc<br />

core Ex9 treatment increases initial lick rate and duration of<br />

licking bursts. These measures are directly correlated with<br />

palatability, so this supports the idea that endogenous GLP-<br />

1R stimulation in NAc reduces the hedonic value of food. We<br />

continue to use other behavioral approaches to discern effects on<br />

palatability and motivation <strong>for</strong> food. Together, our data support<br />

the suggestion that the GLP-1 projection to NAc serves as a link<br />

between gastrointestinal satiation signaling and hedonic aspects<br />

of eating. Acknowledgements: Supported by DK078779<br />

#15 PLATFORM PRESENTATIONS: OLFACTION<br />

Lateral inhibition and odor mixture integration among<br />

distinct subpopulations of olfactory bulb output neurons<br />

imaged in vivo<br />

Michael N Economo, Matt Wachowiak<br />

Brain Institute and Department of Physiology, University of<br />

Utah Salt Lake City, UT, USA<br />

While lateral inhibition between functional processing units<br />

is a fundamental feature of sensory systems, the organization<br />

of lateral inhibitory circuits in the olfactory system and their<br />

impact on odor representations remain unclear. Here, we<br />

addressed this question in the olfactory bulb (OB) using in<br />

vivo two-photon imaging from genetically-defined OB output<br />

neuron populations. We optically recorded activity from mitral<br />

and deep tufted (MT) cells projecting to piri<strong>for</strong>m cortex and<br />

from cholecystokinin-expressing superficial tufted (ST) cells<br />

using Cre-dependent expression of the genetically-encoded<br />

fluorescent calcium indicators GCaMP3 and GCaMP5G. We<br />

compared lateral inhibition to these different populations using<br />

pairs of monomolecular odorants that activated distinct neuron<br />

subsets and their binary odorant mixtures. Both MT and ST<br />

populations showed detectable mixture suppression in a fraction<br />

of neurons (31% <strong>for</strong> MT vs. 36% <strong>for</strong> ST cells), reflecting lateral<br />

inhibition between cell ensembles with distinct odorant response<br />

specificities. However, in these neurons the magnitude of this<br />

suppression was significantly stronger in MT cells vs. ST cells<br />

(70% vs. 25% suppression by the mixture). Thus, these different<br />

OB output populations - which project to distinct cortical targets<br />

- integrate olfactory in<strong>for</strong>mation differently. In addition we found<br />

that odorants often evoked fluorescence decreases in MT somata<br />

and glomerular neuropil in a reliable, odorant-specific manner,<br />

likely reflecting inhibition of spontaneous activity in these<br />

neurons. These experiments thus enable us to, <strong>for</strong> the first time,<br />

directly map the spatial organization of lateral inhibition in vivo<br />

and to characterize how this inhibition varies across different<br />

cell types and OB layers as it shapes early odor representations.<br />

Acknowledgements: NIH NIDCD R01 DC006441 NIH NIDCD<br />

1F32DC012718-01<br />

#16 PLATFORM PRESENTATIONS: OLFACTION<br />

AWnt5 Gradient Patterns the Drosophila Olfactory Map<br />

Huey Hing 1 , Yuping Wu 2 , Lee Fradkin 3 , Jasprina Noordermeer 3<br />

1<br />

SUNY Brockport/Biology Brockport, NY, USA, 2 Stowers Institute <strong>for</strong><br />

Medical Research Kansas City, KS, USA, 3 Leiden University Medical<br />

Center/Molecular Cell Biology Leiden, Netherlands<br />

Although the olfactory map facilitates the encoding of odor<br />

in<strong>for</strong>mation, how its pattern is specified remains largely<br />

unknown. Recent work has shown that the Drosophila olfactory<br />

map is initially <strong>for</strong>med by spatial segregation of the projection<br />

neuron dendrites in the antennal lobe. Here we demonstrate that<br />

the Wnt5 protein specifies projection neuron dendrite positions<br />

prior to arrival of the olfactory receptor axons. Wnt5 is expressed<br />

by a novel set of guidepost cell neurons which are located at the<br />

dorsolateral pole of the antennal lobe and generate a dorsolateral<br />

to ventromedial gradient of Wnt5 in the nascent antennal lobe<br />

neuropil. Loss of wnt5 prevents the appropriate ventral migration<br />

of the projection neuron dendrites while over-expression of<br />

wnt5 severely disrupts dendritic patterning. We also show that<br />

Drl, a known Wnt5 receptor, acts cell-autonomously in the<br />

projection neurons to repress the Wnt5 signal. Decreased drl<br />

function causes projection neuron dendrites to inappropriately<br />

target ventromedially; a defect which is strongly suppressed<br />

by loss of a copy of the wnt5 gene. We propose that the Wnt5-<br />

secreting guidepost cells act to provide positional in<strong>for</strong>mation to<br />

the projection neuron dendrites. Our findings demonstrate the<br />

importance of a Wnt5 gradient in the patterning of the olfactory<br />

map. Acknowledgements: DC010916-01<br />

ORAL ABSTRACTS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

13


#17 PLATFORM PRESENTATIONS: OLFACTION<br />

Sexual dimorphism and experience-dependent plasticity<br />

in mouse vomeronasal neurons<br />

Timothy E Holy, Pei S Xu<br />

Washington University in St. Louis St. Louis, MO, USA<br />

Male and female mice exhibit behaviors particular to their sex,<br />

and these differences presumably reflect sexual dimorphism<br />

in neuronal circuitry. However, in terms of neuronal anatomy<br />

or function in the mouse, there are relatively few reported<br />

differences males and females. Here we asked whether neurons<br />

in the vomeronasal organ, a social odor- and pheromone-sensing<br />

neuroepithelium, differed functionally between males and<br />

females. We per<strong>for</strong>med exhaustive high-speed calcium imaging<br />

from intact vomeronasal epithelia in 26 imaging volumes from<br />

male and female mice, measuring the responses of more than a<br />

quarter-million individual sensory neurons. Using a battery of<br />

sulfated steroids, a class of odors originally isolated from mouse<br />

urine, we found that the large majority of responsive neuronal<br />

types were found in equal abundance in both males and females.<br />

However, we found restricted cases of clear sexual dimorphism,<br />

including one neuronal type that was more than one-hundredfold<br />

more common in males than in females, by far the<br />

strongest dimorphism ever reported in the mammalian central<br />

nervous system. We then explored the mechanism generating<br />

this dimorphism. Surprisingly, male/female differences<br />

depended entirely on the history of sensory experience, as<br />

vomeronasal organs from males could be converted to a pattern<br />

indistinguishable from females simply by prolonged exposure<br />

to the odors of female mice. The finding that a strong neuronal<br />

dimorphism is determined entirely by experience, in a sensory<br />

system long believed to be devoted to “innate” responses,<br />

raises important new questions about the roles of “nature”<br />

and “nurture” in brain architecture. Acknowledgements:<br />

NIH-NIDCD R01 DC005964 NIH-NINDS/NIAAA R01<br />

NS068409 NIH DP1 OD006437<br />

#18 PLATFORM PRESENTATIONS: OLFACTION<br />

Rein<strong>for</strong>cement of Sexual Attraction through<br />

Pheromone-Induced Associative Learning<br />

Jane L Hurst, Sarah A Roberts, Emma Hoffman,<br />

Amanda J Davidson, Lynn McLean, Robert J Beynon<br />

University of Liverpool / Institute of Integrative Biology Liverpool,<br />

United Kingdom<br />

Scents play an integral role in mediating reproductive<br />

interactions in many species, allowing animals to recognize<br />

and locate individual conspecifics of the opposite sex, assess<br />

their attractiveness and stimulate mating. The urine used by<br />

male house mice to advertise their location and competitive<br />

dominance contains many androgen-dependent volatiles together<br />

with a high concentration of major urinary proteins (MUPs) that<br />

bind and slow the release of these volatiles. Females that detect<br />

airborne urinary volatiles are attracted to approach and sniff<br />

the scent marks closely, but it is contact with an atypical malespecific<br />

MUP named darcin that reliably elicits female sexual<br />

attraction to spend time near male scent. Darcin will elicit this<br />

attraction even in the absence of all other urinary components,<br />

while male urine without darcin is no more attractive than<br />

female urine. Importantly, though, darcin acts not only through<br />

direct attraction to spend time near the pheromone itself. It is<br />

also highly potent in stimulating associative learning such that<br />

animals rapidly learn the same attraction towards the individual<br />

airborne odor detected in association with darcin, targeting<br />

attraction to a specific male. Darcin also rapidly conditions<br />

preference <strong>for</strong> spatial cues associated with its location, such that<br />

mice relocate and prefer to spend time in the site even when scent<br />

is absent. This conditioned place preference is remembered <strong>for</strong><br />

approximately 14 days. Preference <strong>for</strong> multiple locations can be<br />

remembered, while the relative amount of darcin in competing<br />

male scents influences female conditioned preferences. This<br />

reveals that pheromone-induced social learning can both<br />

target and strongly rein<strong>for</strong>ce female sexual attraction towards<br />

individual males even though male mice all produce the same sex<br />

pheromone. Acknowledgements: These studies were supported<br />

by the Biotechnology and Biological Science Research Council<br />

(BBC503897 and BB/J002631/1) and the Natural Environment<br />

Research Council (NE/G018650), UK<br />

#19 PLATFORM PRESENTATIONS: OLFACTION<br />

Behavioral effects of bulbar neuromodulation<br />

Christiane Linster, Sasha Devore, T Samuel Dillon, Olga Escanilla,<br />

Matthew Lewis, Laura Manella<br />

Cornell University Neurobiology and Behavior Ithaca, NY, USA<br />

The olfactory bulb (OB) is modulated by a number of extrinsic<br />

and intrinsic modulatory substances. We here synthesize work<br />

from our lab investigating the effects of cholinergic (ACh),<br />

noradrenergic (NE), serotonergic (5HT), dopaminergic (DA) and<br />

hormonal modulation on olfactory perception. We compare the<br />

behavioral effects by using well–established behavioral paradigms<br />

such as olfactory habituation memory and reward-association.<br />

We find that ACh modulation affects the discrimination of<br />

chemically and perceptually similar odorants (Mandairon et al.<br />

2006; Chaudhury et al. 2009), with specific roles <strong>for</strong> muscarinic<br />

and nicotinic receptors (Devore et al., in prep). NE, while also<br />

modulating odor discrimination of chemically and perceptually<br />

similar odorants (Mandairon et al. 2006; see also Doucette et<br />

al. 2007), also plays a specific role in regulating signal to noise<br />

ratio when very low concentration odorants are used (Escanilla<br />

et al 2010). In a direct comparison, NE, but not ACh, is shown<br />

to be functionally important during learning of very low<br />

concentration odorants (Escanilla et al 2012). DA, intrinsic to<br />

the bulb, modulated odor concentration perception via activation<br />

of D2 but not D1 receptors (Escanilla et al. 2009). Blockade of<br />

5HT receptors impaired habituation memory <strong>for</strong>mation and<br />

specificity, as well as reward-driven discrimination at low, but<br />

not higher odor concentrations (Lewis and Linster, in prep).<br />

Both hormonal (estrodial, in mice) and NE manipulations (in<br />

rats) extended the duration of an olfactory habituation memory<br />

(Dillon et al, in prep; Manella et al. in prep). In summary, the<br />

behavioral effects of OB neuromodulators, while similar at first<br />

ORAL ABSTRACTS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

14


sight, can be dissociated to ascribe specific roles to each of these<br />

substances when task demands are manipulated in different<br />

ways. Acknowledgements: NIH/NIDCD RO1DC009948 NIH/<br />

NIDCD RO1DC008701 NIH/NIDCD F32DC011974 l’Oreal<br />

USA fellowship <strong>for</strong> women in science<br />

#21 PLATFORM PRESENTATIONS: OLFACTION<br />

The Scent of a Human: A Mosquito Olfactory Receptor<br />

Neuron that Mediates Attraction to Human Skin Odor<br />

Genevieve M. Tauxe, Anandasankar Ray<br />

Department of Entomology, University of Cali<strong>for</strong>nia, Riverside<br />

Riverside, CA, USA<br />

#20 PLATFORM PRESENTATIONS: OLFACTION<br />

Olfactory Dysfunction Predicts 5-year Mortality in<br />

Older Adults<br />

Jayant M. Pinto 1 , Kristen E. Wroblewski 2 , David W. Kern 3 ,<br />

L. Philip Schumm 2 , Martha K. McClintock 4<br />

1<br />

Section of Otolaryngology-Head and Neck Surgery, The University<br />

of Chicago Chicago, IL, USA, 2 Department of Health Studies,<br />

The University of Chicago Chicago, IL, USA, 3 Department of<br />

Comparative Human, The University of Chicago Chicago, IL, USA,<br />

4<br />

The Institute <strong>for</strong> Mind and Biology, The University of Chicago<br />

Chicago, IL, USA<br />

We sought to determine if loss of olfactory function among<br />

older US adults is a harbinger of overall health decline<br />

predicting 5-year mortality. In 2005-6 The National Social Life,<br />

Health and Aging Project (NSHAP) interviewed 3,005 adults<br />

aged 57-85 (Wave 1) representative of the diverse aging US<br />

population. Olfactory function was measured with 5 Sniffin’<br />

Sticks pens and word/picture prompts to classify participants<br />

as normosmic (4-5 correct), hyposmic (2-3 correct), or anosmic<br />

(0-1 correct). Five years later NSHAP Wave 2 established<br />

participant mortality. Logistic regression used olfactory function<br />

at Wave 1 to estimate the odds of death at Wave 2, controlling<br />

<strong>for</strong> baseline factors known to influence both olfaction and<br />

mortality (age, gender, education, race/ethnicity, cognitive<br />

function, physical and mental health, and tobacco/alcohol use).<br />

Olfactory identification at Wave 1 strongly predicted subsequent<br />

5-year all-cause mortality: fully 39% of anosmic subjects had<br />

died compared to 19% in the hyposmic group and 10% in the<br />

normosmic group (p


maintenance; demonstrate a contribution of Shh responding,<br />

Gli1 labeled progeny to taste bud cells and cells of the papilla;<br />

and present distinctive effects of activating Shh transcription<br />

factors in adult tongue. Defined proliferation niches of active<br />

Shh signaling suggest that epithelium and mesenchyme harbor<br />

multiple stem and progenitor cell compartments. In sum, Shh<br />

has roles to <strong>for</strong>m and maintain fungi<strong>for</strong>m papillae and taste<br />

buds, most likely via stage-specific autocrine and/or paracrine<br />

signaling, with epithelial/mesenchymal interactions. Neural<br />

crest cells also contribute to developing and postnatal lingual<br />

epithelium and mesenchyme, including taste papillae and taste<br />

buds. Further, Shh signaling in neural crest cells participates in<br />

patterning the tongue. In all, cells originating in several lingual<br />

tissue areas contribute to the stem and progenitor compartments<br />

that are active in <strong>for</strong>ming and maintaining taste buds and<br />

papillae. Supported by NIH NIDCD Grant DC000456.<br />

#24 SYMPOSIUM:<br />

STEM AND PROGENITOR CELLS FOR<br />

TASTE BUDS — DEVELOPMENT AND RENEWAL<br />

In search of adult taste stem cells<br />

Karen Yee 1 , Yan Li 1 , Kevin Redding 1 , Ken Iwatsuki 2 ,<br />

Robert Margolskee 1 , Peihua Jiang 1<br />

1<br />

Monell Chemical Senses Center, 3500 Market Street, Philadelphia,<br />

PA, USA, 2 Institute <strong>for</strong> Innovation, Ajinomoto Co., Inc.,<br />

Kawasaki-ku, Kawasaki, Japan<br />

Recently, a great deal of progress has been made in identifying<br />

reliable markers <strong>for</strong> adult stem cells <strong>for</strong> many regenerative<br />

mammalian tissues. For instance, Lgr5 (leucine-rich repeatcontaining<br />

G-protein coupled receptor 5) is a bona fide marker<br />

<strong>for</strong> adult stem cells in intestine, stomach, and hair follicle. In<br />

the small intestine of mice the Polycomb group protein Bmi1<br />

marks another population of stem cells distinct from the Lgr5+<br />

stem cells. Taste epithelium also regenerates constantly, yet the<br />

identity of adult taste stem cells remains elusive. In this study we<br />

set out to determine if Lgr5 and Bmi1 mark adult taste stem/<br />

progenitor cells. We found that Lgr5 is strongly expressed in<br />

cells at the bottom of trench areas at the base of circumvallate<br />

and foliate taste papillae and weakly expressed in the basal area<br />

of taste buds and that Lgr5-expressing cells in posterior tongue<br />

are a subset of K14-positive epithelial cells. Lineage-tracing<br />

experiments using an inducible Lgr5-Cre knock-in allele in<br />

combination with Rosa26-LacZ and Rosa26-tdTomato reporter<br />

strains showed that Lgr5-expressing cells gave rise to taste cells,<br />

perigemmal cells, along with self-renewing cells at the bottom<br />

of trench areas at the base of circumvallate and foliate papillae.<br />

Moreover, using subtype-specific taste markers, we found that<br />

Lgr5-expressing cell progeny include all three major types of<br />

adult taste cells. Our results indicate that Lgr5 may mark adult<br />

taste stem cells in the posterior portion of the tongue. In contrast,<br />

our lineage tracing experiments using Bmi1-Cre; Rosa26-<br />

LacZ showed that Bmi1 does not mark adult taste stem cells.<br />

Acknowledgements: This work was supported by NIH grants<br />

DC0101842 (P.J.), DK081421 (R.F.M.), DC003055 (R.F.M.),<br />

P30DC011735 (R.F.M) and a grant from the Commonwealth of<br />

Pennsylvania Department of Health (P.J.)<br />

#25 SYMPOSIUM:<br />

STEM AND PROGENITOR CELLS FOR<br />

TASTE BUDS — DEVELOPMENT AND RENEWAL<br />

Cell types in adult taste buds: distinct longevities and origins<br />

Nirupa Chaudhari<br />

Department of Physiology and Biophysics, and Program in<br />

Neurosciences, University of Miami Miller School of Medicine,<br />

Miami, FL, USA<br />

Mammalian taste buds are repopulated throughout adult life<br />

with new cells that are born in the basal epithelium immediately<br />

surrounding them. Taste buds contain several molecularly and<br />

functionally distinct classes of cells, but it is unclear whether<br />

each class is derived from a separate progenitor cell pool, and<br />

has similar longevity. Using high resolution confocal microscopy,<br />

and 4-color fluorescent labeling, we examined whether the<br />

previously described average life-time of taste bud cells applies to<br />

each of the cell types that make up the taste bud. After a single<br />

EdU injection to label newly born cells, we fitted exponential<br />

decay curves to the disappearance of labeled nuclei from each<br />

cell type. While sweet-, bitter- and umami-sensing Type 2 cells<br />

displayed a half-life of ≈8 days, sour-sensing neuron-like Type<br />

3 cells were much longer-lived, with a half-life of over 22 days.<br />

Curiously, many post-mitotic cells had a prolonged quiescence<br />

inside taste buds be<strong>for</strong>e differentiating into mature taste cells.<br />

We also evaluated whether all taste cell types arise from a<br />

common pool of progenitor cells using lineage-tracing analyses<br />

in adult KERATIN14-cre/ERT;Rosa26-YFP mice. Non-taste<br />

keratinocytes were produced rapidly (within 2 days) from K14+<br />

progenitors; Type I glial-like cells became YFP+ ≈10 days after<br />

induction, whereas Type II cells contained YFP+ cells only after<br />

≈20 days. In stark contrast, Type III cells did not noticeably<br />

acquire the YFP lineage-label even 60 days after induction. Thus,<br />

while most cells of the taste bud arise from K14+ progenitors,<br />

the dynamics of their appearance suggests that several separate<br />

progenitor cell pools replenish adult taste buds during normal<br />

renewal. Supported by NIH/NIDCD R01DC6308<br />

#26 SYMPOSIUM:<br />

STEM AND PROGENITOR CELLS FOR<br />

TASTE BUDS — DEVELOPMENT AND RENEWAL<br />

Development and regeneration of the inner ear: Control of<br />

cell division and differentiation of sensory progenitors<br />

Neil Segil<br />

Division of Cell Biology and Genetics, House Research Institute;<br />

and Department of Cell and Neurobiology, University of<br />

Southern Cali<strong>for</strong>nia.<br />

Loss of the sensory hair cells of the inner ear is the major cause<br />

of deafness and balance disorders. Hair cells are extremely<br />

sensitive to various environmental stressors such as ototoxic<br />

drugs noise, and aging. In mammals, the failure to regenerate<br />

these cells is complete, making hair cell loss a major global health<br />

problem. In contrast to the failure of regeneration in mammals,<br />

non-mammalian vertebrates can efficiently regenerate sensory<br />

hair cells through a combination of direct transdifferentiation<br />

ORAL ABSTRACTS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

16


of the surrounding supporting cells, as well as by stimulated<br />

cell cycle reentry and subsequent differentiation of supporting<br />

cells into new functional hair cells and supporting cells.<br />

In spite of the failure of hair cell regeneration in mammals,<br />

recent work in the mouse has shown a latent regenerative<br />

potential <strong>for</strong> supporting cells to act as progenitors by both<br />

cell cycle reentry and transdifferentiation during the perinatal<br />

period, a potential that appears to be rapidly lost during early<br />

postnatal maturation. In this talk I will present recent findings<br />

on the molecular mechanisms that govern cell cycle reentry of<br />

otherwise postmitotic supporting cells, as well as mechanisms<br />

governing cell fate decisions between sensory hair cells, and the<br />

various supporting cell types in the developing and perinatal<br />

organ of Corti. I will also discuss the possibilities <strong>for</strong> future<br />

manipulation of supporting cells <strong>for</strong> the purpose of regeneration<br />

of lost sensory hair cells.<br />

#27 SYMPOSIUM:<br />

STEM AND PROGENITOR CELLS FOR<br />

TASTE BUDS — DEVELOPMENT AND RENEWAL<br />

Cellular basis of taste dysfunction following head and<br />

neck irradiation<br />

Linda Barlow 1,2 , Alon Bajayo 1,2 , Ha Nguyen 1,2,3 , Mary Reyland 4<br />

1<br />

Department of Cell & Developmental Biology, University of Colorado<br />

School of Medicine, Aurora CO 80045, USA, 2 the Rocky Mountain<br />

Taste & Smell Center (RMTSC), University of Colorado School<br />

of Medicine, Aurora CO 80045, USA, 3 Hanoi Medical University,<br />

Department of Histology and Embryology, No. 1 Ton That Tung,<br />

Hanoi, Vietnam , 4 Dept of Craniofacial Biology, University of Colorado<br />

School of Dental Medicine, Aurora CO 80045, USA<br />

Taste dysfunction frequently occurs following radiotherapy<br />

<strong>for</strong> head and neck cancer. Importantly, patients with reduced<br />

taste function tend to lack appetite and eat far less, leading to<br />

weight loss and a significantly compromised quality of life. To<br />

determine the cellular targets contributing to taste dysfunction,<br />

we developed a mouse model of head and neck irradiation.<br />

We find that proliferating taste bud progenitor cells are<br />

immediate and direct targets of radiation damage. Specifically,<br />

head and neck irradiation results in: 1) increased apoptosis of<br />

progenitors within 24 hours of radiation exposure; and 2) a<br />

transient cessation in proliferation, lasting ~3 days. This latter<br />

effect results in an interrupted supply of new postmitotic taste<br />

cells to buds, and accounts <strong>for</strong> the subsequent reduction in<br />

differentiated taste cells seen at 1 week post-irradiation. We are<br />

now investigating the role of the novel protein kinase C delta<br />

iso<strong>for</strong>m (PKCδ) in irradiation-induced taste epithelial injury.<br />

PKCδ is a key regulator of irradiation-induced apoptosis, and<br />

suppression of PKCδ in vitro and genetic loss in vivo protects<br />

salivary gland cells from cell death. PKCδ is also implicated in<br />

maintenance of cell cycle arrest required <strong>for</strong> DNA repair in UV<br />

damaged human keratinocytes. Thus, we hypothesize that loss<br />

of PKCδ may protect taste progenitor cells from death, and/<br />

or promote their continued mitosis following irradiation injury.<br />

This model is supported by our pilot data, which suggest that,<br />

while PKCδ -/- mice possess normal taste epithelia, in response<br />

to irradiation, the taste bud progenitor population continues to<br />

proliferate. Thus, our studies point to PKCδ as a potential future<br />

target <strong>for</strong> interventional treatment <strong>for</strong> taste loss in head and neck<br />

cancer patients treated with radiotherapy. Supported by NIH/<br />

NIDCD R21DC011713 to LAB and MER, and P30DC004657<br />

to D. Restrepo.<br />

#28 PLATFORM PRESENTATIONS —<br />

POLAK YOUNG INVESTIGATOR AWARD WINNERS<br />

Trace amine-associated receptors mediate behavioral<br />

aversion in mice<br />

Adam Dewan 1 , Rodrigo Pacifico 1 , Dmitry Rinberg 2 , Thomas Bozza 1<br />

1<br />

Department of Neurobiology, Northwestern University Evanston, IL,<br />

USA, 2 Neuroscience Institute, New York University Langone Medical<br />

Center New York, NY, USA<br />

The Trace Amine-Associated Receptors (TAARs) are a small<br />

set of evolutionarily conserved main olfactory receptors whose<br />

contribution to chemosensory function is not known. Our<br />

previous data show that a majority of the TAARs are mapped to<br />

a discrete group of highly sensitive amine-responsive glomeruli<br />

in the dorsal olfactory bulb of the mouse. Amines have been<br />

implicated as social cues and/or predator-derived chemosignals<br />

in rodents. We have used a combination of behavior and in<br />

vivo optical imaging to examine the functional consequences<br />

of genetically removing TAAR genes in mice. We find that<br />

deleting all 14 olfactory TAARs abolishes high sensitivity amine<br />

responses in the dorsal bulb and eliminates aversion that mice<br />

display to structurally diverse amines and to the volatiles of<br />

predator cat urine. Moreover, removing a single TAAR gene<br />

(Taar4) produces an odor-specific deficit in sensitivity and<br />

abolishes behavioral aversion to phenylethylamine—a chemical<br />

that is enriched in predator cat urine. Our data reveal that the<br />

TAARs mediate aversive responses in some behavioral contexts,<br />

and that individual TAAR genes contribute significantly to<br />

amine perception in mice. Acknowledgements: This work was<br />

supported by grants from NIH/NIDCD (R01DC009640), The<br />

Whitehall Foundation, and The Brain Research Foundation.<br />

#29 PLATFORM PRESENTATIONS —<br />

POLAK YOUNG INVESTIGATOR AWARD WINNERS<br />

Excess Wnt/b-catenin signaling in lingual epithelial<br />

progenitors drives production of type I taste cells at the<br />

expense of all other lingual epithelial cell fates<br />

Dany Gaillard 1 , Sarah E Millar 2 , Fei Liu 3 , Linda A Barlow 1<br />

1<br />

University of Colorado Anschutz Medical Campus, Department<br />

of Cell & Developmental Biology and the Rocky Mountain Taste &<br />

Smell Center Aurora, CO, USA, 2 University of Pennsylvania School<br />

of Medicine, Departments of Dermatology and Cell & Developmental<br />

Biology Philadelphia, PA, USA, 3 Institute <strong>for</strong> Regenerative Medicine at<br />

Scott & White Hospital, Texas A&M University System Health Science<br />

Center Temple, TX, USA<br />

In adult mice, taste cells are continually renewed from Keratin<br />

(K)14+ basal keratinocytes. As a population, K14+ basal cells<br />

ORAL ABSTRACTS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

17


self renew, as well as produce post-mitotic cells which either<br />

differentiate into lingual epithelial cells (K13+), or enter buds<br />

and differentiate into type I, II and III taste cells (K8+). We<br />

previously showed that Wnt/b-catenin signaling is active in<br />

cells in and around taste buds of adult mice (Gaillard & Barlow,<br />

2011), suggesting that this pathway may regulate several aspects<br />

of taste cell renewal. To test this idea, we used inducible Cre-lox<br />

technology to drive b-catenin gain of function (GOF) in K14+<br />

basal keratinocytes throughout the tongue epithelium, including<br />

the fungi<strong>for</strong>m and circumvallate papillae (CVP). In the CVP<br />

trenches, K13+ cells located between taste buds vanished in the<br />

GOF, and instead all cells within the CVP epithelium expressed<br />

K8. Using immunomarkers <strong>for</strong> each of the 3 taste cell types,<br />

we found that this expanded taste CVP epithelium comprised<br />

primarily NTPdase2+ type I cells, with little or no change in<br />

the numbers of type II and III cells. Likewise, in the anterior<br />

tongue, b-catenin GOF induced multiple K8+ cell clusters within<br />

fungi<strong>for</strong>m papillae, as well as numerous ectopic K8+ cell clusters<br />

in non-taste epithelium; all of these cell clusters were exclusively<br />

NTPdase2+, and were devoid of expression of markers <strong>for</strong> type<br />

II and III taste cells. Our data indicate that excess Wnt/b-catenin<br />

signaling drives epithelial progenitors to produce daughter cells<br />

committed to a taste fate (K8+) at the expense of a non-taste<br />

fate (K13+), and moreover constrains these newly generated<br />

taste cells to a type I cell fate. We are now examining what<br />

cellular mechanisms are triggered by excess b-catenin, and how<br />

these changes in cell renewal result in the GOF phenotype.<br />

Acknowledgements: Supported by an American Heart<br />

<strong>Association</strong> fellowship to DG, NIH/NIDCD DC008373 and<br />

DC012383 to LAB, and DC004657 to D. Restrepo.<br />

#30 PLATFORM PRESENTATIONS —<br />

POLAK YOUNG INVESTIGATOR AWARD WINNERS<br />

Target-defined olfactory bulb output streams isolated using<br />

retrograde infection with recombinant viral vectors<br />

Markus Rothermel, Christine Zabawa, Daniela Brunert,<br />

Marta Diaz-Quesada , Matt Wachowiak<br />

Brain Institute and Department of Physiology Salt Lake City, UT, USA<br />

Recombinant viral vectors are an attractive tool <strong>for</strong> cell-type<br />

specific transgene expression, especially when Cre-dependent<br />

vectors are combined with Cre-expressing mouse lines.<br />

However, cell-type specific promoters do not always yield<br />

sufficient specificity <strong>for</strong> isolating functionally distinct neuronal<br />

populations. In the olfactory system mitral and tufted cells (MT)<br />

of the olfactory bulb (OB) constitute a heterogenous population<br />

with distinct dendritic organization, response properties and<br />

projections to olfactory cortex. Here, we demonstrate that by<br />

combining viral tools with retrograde infection via their axonal<br />

processes, MT cells can be defined by projection target. We<br />

injected Cre-dependent AAV vectors into various regions of<br />

olfactory cortex in Cdhr1-cre mice. Virus injection into anterior<br />

piri<strong>for</strong>m cortex (PC) led to robust and widespread transgene<br />

expression in MT cells throughout the OB. To establish that<br />

expression patterns were due to retrograde infection we targeted<br />

additional olfactory cortical areas: injection into posterior PC<br />

or posterior cortical amygdala led to expression exclusively in<br />

mitral cells with lateral dendrites in the deep external plexi<strong>for</strong>m<br />

layer, while injection into medial amygdala led to expression<br />

solely in mitral cells of the accessory olfactory bulb. Retrograde<br />

infection was effective using multiple transgenes including<br />

GCaMP and ChR2, allowing <strong>for</strong> optical imaging, optical control<br />

or optically-assisted electrophysiology of distinct OB output<br />

streams defined by their projection target. Retrograde infection<br />

was also effective <strong>for</strong> projection neurons in other brain regions.<br />

These results establish a valuable, easily-used tool <strong>for</strong> achieving<br />

combinatorial specificity in transgene expression to monitor<br />

and manipulate precisely-defined neuron populations in vivo.<br />

Acknowledgements: Supported by DFG and NIDCD<br />

#31 PLATFORM PRESENTATIONS —<br />

POLAK YOUNG INVESTIGATOR AWARD WINNERS<br />

Major contribution of Gao-dependent vomeronasal<br />

chemoreception to sexual and reproductive behavior in<br />

female mice<br />

Livio Oboti 1 , Trese Leinders-Zufall 1 , Eric Jacobi 1,3 , Lutz Birnbaumer 2 ,<br />

Frank Zufall 1 , Pablo Chamero 1<br />

1<br />

Department of Physiology, University of Saarland School of<br />

Medicine Homburg, Germany, 2 Laboratory of Neurobiology, Division<br />

of Intramural Research, National Institutes of Health Research<br />

Triangle Park, NC, USA, 3 Deutsches Zentrum für Neurodegenerative<br />

Erkrankungen (DZNE) Heidelberg, Germany<br />

Optimal reproductive fitness is essential <strong>for</strong> the biological<br />

success and survival of species. The vomeronasal organ (VNO)<br />

is strongly implicated in the display of sexual and reproductive<br />

behaviors in female mice, yet the role that apical and basal<br />

vomeronasal neuron populations play in controlling these<br />

gender-specific behaviors remain largely unclear. To dissect<br />

neural pathways underlying these functions, we genetically<br />

inactivated the basal VNO layer using conditional, cell-specific<br />

ablation of the G protein Gao. Female mice mutant <strong>for</strong> Gao<br />

show severe alterations in mate recognition, mating, and<br />

reproduction. Male pheromonal cues fail to accelerate puberty<br />

onset and estrous synchronization in these mice. Gao mutant<br />

females exhibit a striking reduction in sexual receptivity or<br />

lordosis behavior to males, but gender discrimination seems<br />

to be intact. These mice also show a loss in scent ownership<br />

recognition that requires a learned association with a nonvolatile<br />

ownership signal contained in the high molecular weight fraction<br />

of urine, and they show high pregnancy failure rates in the<br />

Bruce effect assay. These results indicate that sensory neurons<br />

of the Gao-expressing vomeronasal subsystem, together with<br />

the receptors they express and the molecular cues they detect,<br />

control a diverse range of fundamental mating and reproductive<br />

behaviors in female mice. Acknowledgements: This work was<br />

supported by grants from the Deutsche Forschungsgemeinschaft<br />

to P.C. (CH 920/2-1), F.Z. (SFB 894) and T.L-Z. (SFB 894),<br />

the Intramural Research Program of the NIH to L.B. (Project<br />

Z01 ES-101643), and the Volkswagen Foundation (to T.L.-Z.).<br />

E.J. was supported by the DFG-funded International Graduate<br />

Program GK1326. T.L.-Z. is a Lichtenberg Professor of the<br />

Volkswagen Foundation.<br />

ORAL ABSTRACTS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

18


#32 PLATFORM PRESENTATIONS —<br />

POLAK YOUNG INVESTIGATOR AWARD WINNERS<br />

Congruency matters: dual cortical processing of<br />

visual-olfactory integration<br />

Kathrin Ohla 1,2 , Johan N Lundström 2,3,4<br />

1<br />

German Institute of Human Nutrition Potsdam-Rehbrücke Nuthetal,<br />

Germany, 2 Monell Chemical Senses Center Philadelphia, PA, USA,<br />

3<br />

Karolinska Institute Stockholm, Sweden, 4 University of Pennsylvania<br />

Philadelphia, PA, USA<br />

Be<strong>for</strong>e ingestion, the visual appearance and odor of a food<br />

constitute its primary sensory inputs. Whether these distinct<br />

sensory events are perceived as one entity shapes their impact<br />

on subsequent food choice and possibly also food perception.<br />

We hypothesized that the degree of perceived concordance<br />

of the sensory inputs influences multisensory integration<br />

processes. To date, the behavioral consequences and brain<br />

mechanisms underlying these processes are poorly understood<br />

and were investigated with the present study. We used electrical<br />

neuroimaging analyses of the electroencephalographic (EEG)<br />

responses following olfactory-visual stimulation in humans.<br />

Stimuli were odor-image pairs presented as 100% congruent,<br />

50% congruent and 100% incongruent. Participants rated the<br />

stimuli <strong>for</strong> congruence, pleasantness and intensity. As expected,<br />

concordant stimulus pairs were perceived as more congruent<br />

and as more pleasant than mixed and incongruent pairs.<br />

Wave<strong>for</strong>m analysis yielded significant amplitude augmentation<br />

<strong>for</strong> congruent as compared to incongruent odor-image pairs<br />

between 120-200ms post stimulus onset. Source analysis revealed<br />

that the activation differences origin in visual cortex and left<br />

inferior temporal gyrus. Later differences were observed between<br />

400-700ms in the parietal lobe, lateral frontal cortex and the<br />

bilateral insula. Concordance between olfactory-visual stimuli<br />

was associated with increased pleasantness and activation in<br />

unimodal visual and olfactory areas as well as in multimodal<br />

areas. The results suggest that cross-modal integration is a<br />

dynamic process regulated by both unisensory as well as higher<br />

order integration areas and that this process is modulated by<br />

learned associations and perceived congruence between the<br />

sensory inputs.<br />

#33 PLATFORM PRESENTATIONS —<br />

POLAK YOUNG INVESTIGATOR AWARD WINNERS<br />

Processing of hedonic and chemosensory features of taste<br />

in medial prefrontal and gustatory cortices<br />

Ahmad Jezzini, Luca Mazzucato, Giancarlo La Camera,<br />

Alfredo Fontanini<br />

Dept of Neurobiology & Behavior, SUNY Stony Brook, NY, USA<br />

The gustatory cortex (GC) is the main cortical recipient of<br />

taste-related signals, however it is not the only cortical area<br />

involved in processing taste. Upon elaborating gustatory signals,<br />

GC sends in<strong>for</strong>mation to the orbitofrontal cortex (OFC) and<br />

the medial prefrontal cortex (mPFC). While considerable<br />

amount of work has been devoted to understanding how GC<br />

and OFC process different features of a gustatory experience,<br />

less is known regarding the role of mPFC. We investigated<br />

the involvement of mPFC in taste processing by comparing<br />

its responses to gustatory stimuli with those observed in GC.<br />

Eight rats were chronically implanted with movable bundles<br />

of electrodes in the ipsilateral mPFC and GC. Extracellular<br />

recordings were per<strong>for</strong>med while 4 tastes were passively delivered<br />

through intraoral cannulae. The results showed significant<br />

taste related activity in mPFC. In comparison to GC, mPFC<br />

was less responsive to taste (49% of mPFC units responded to<br />

taste vs 75% in GC) and firing rates were lower in mPFC than<br />

GC. While taste selectivity was more pronounced in GC than<br />

mPFC, activity in mPFC appeared more strongly modulated by<br />

palatability. Results from a classification analysis revealed that<br />

units in GC decode taste quality more successfully than mPFC,<br />

and mPFC units encode tastes according to their palatability.<br />

Analysis of the time course of responses further revealed that,<br />

contrary to the GC where palatability coding occurs within<br />

the first 2 seconds following taste delivery, palatability coding<br />

in mPFC lasts <strong>for</strong> as long as 5 seconds. Furthermore, analysis<br />

of firing rates evoked by tastes revealed a bias toward aversive<br />

stimuli in mPFC. Altogether our results suggest a role of<br />

the mPFC in taste processing and more specifically point to<br />

mPFC as a cortical area involved in coding of palatability.<br />

Acknowledgements: NIDCD Grant R01-DC010389<br />

#34 SYMPOSIUM:<br />

NEW APPROACHES TO PHYSIOLOGY<br />

AND BEHAVIOR IN AWAKE RODENTS<br />

New approaches to physiology and behavior in awake rodents<br />

Stephen Shea 1 , Dinu Albeanu 1 , Alfredo Fontanini 2 ,<br />

Venkatesh Murthy 3 , Andreas Schaefer 4 , Ben Strowbridge 5<br />

1<br />

Cold Spring Harbor Laboratory Cold Spring Harbor, NY, USA,<br />

2<br />

Stony Brook University Stony Brook, NY, USA, 3 Harvard University<br />

Cambridge, MA, USA, 4 MPI Heidelberg Heidelberg, Germany, 5 Case<br />

Western Reserve University Cleveland, OH, USA<br />

As our understanding of neural circuitry grows more<br />

sophisticated, there is increasing interest in studying neuronal<br />

processing during volitional behavior in awake animals. At the<br />

same time, recent years have seen the emergence of many new<br />

techniques that allow unprecedented access to observe and<br />

control neural activity with spatiotemporal precision. These<br />

sensitive techniques can be difficult to implement and chemical<br />

stimulus control can be problematic in freely behaving animals.<br />

For these reasons, many labs are developing hybrid approaches<br />

in head-fixed preparations that fuse these new tools with rich<br />

behavioral paradigms in awake, behaving animals. The goal<br />

of this symposium will be to bring together investigators who<br />

are applying high resolution tools in awake animals to break<br />

new ground in our appreciation of the state modulation of<br />

chemosensory circuits and their governance of behavior.<br />

ORAL ABSTRACTS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

19


#35 SYMPOSIUM:<br />

NEW APPROACHES TO PHYSIOLOGY<br />

AND BEHAVIOR IN AWAKE RODENTS<br />

Dramatic state-dependency of the activity of granule cells in<br />

the mouse main olfactory bulb<br />

Stephen D Shea, Brittany N Cazakoff, Kerensa L Crump, Billy Y Lau<br />

Cold Spring Harbor Laboratory Cold Spring Harbor, NY, USA<br />

It is becoming increasingly clear that olfactory representations in<br />

the main olfactory bulb (MOB) are substantially re<strong>for</strong>matted in<br />

awake rodents. In addition to elevated rates and altered temporal<br />

structure in the spike discharge of mitral/tufted cells (MT),<br />

the wakeful state is marked by a fusion of sensory input-driven<br />

responses with activity that reflects attention, experience, and<br />

behavioral task contingencies. It seems likely that as a key target<br />

of many neuromodulatory systems and corticofugal feedback<br />

pathways, the extensive network of inhibitory MOB granule<br />

cells (GC) is instrumental in the state-dependent sculpting of<br />

MT activity patterns. Despite this predicted critical role, few<br />

published studies have demonstrably made electrophysiological<br />

recordings from these small cells. Moreover, none have been<br />

reported in awake animals. As a first step towards closing this<br />

gap, we recently developed reliable methods <strong>for</strong> recording and<br />

labeling GC in awake, head-fixed mice. Our preparation allows<br />

us to directly compare the activity and sensory responses of<br />

the same GC during wakefulness and inhalant anesthesia.<br />

Our data reveal that GC in awake mice are dramatically more<br />

spontaneously active, and exhibit stronger, more broadly-tuned<br />

sensory responses that include both increases and decreases<br />

in spike rate. Under either of two pharmacologically-distinct<br />

anesthetics, many of these cells emit very few spontaneous<br />

or stimulus-driven spikes. Those that have somewhat higher<br />

spontaneous rates still exhibit little response to odors. We are<br />

currently quantifying the variable respiratory coupling of GC in<br />

awake animals, and assessing the effects of stimulus novelty or<br />

familiarity on GC odor responses.<br />

#36 SYMPOSIUM:<br />

NEW APPROACHES TO PHYSIOLOGY<br />

AND BEHAVIOR IN AWAKE RODENTS<br />

Response properties of cortico-bulbar feedback and granule<br />

cells in awake head-fixed mice<br />

Dinu F Albeanu, Hong goo Chae, Gonzalo H Otazu<br />

Cold Spring Harbor Laboratory Cold Spring Harbor, NY, USA<br />

Sensory circuits integrate inputs from the environment, as well<br />

as feedback signals from higher brain regions, in a close loop<br />

manner. The interplay of feed-<strong>for</strong>ward and feedback signals<br />

has been proposed to be fundamental <strong>for</strong> learning and memory<br />

recall. Though rich cortical feedback projections innervate<br />

the mouse olfactory bulb, to date, little is known about their<br />

contribution to olfactory processing. Cortico-bulbar feedback<br />

primarily targets the granule cells (GC), which <strong>for</strong>m extensive<br />

dendro-dendritic synapses with the mitral/tufted cells. To<br />

examine how cortical feedback shapes processing in the bulb,<br />

we used genetically encoded calcium indicators (GCaMP3&5)<br />

and multiphoton imaging to monitor the odor evoked responses<br />

of feedback fibers and granule cells, in awake head-fixed mice.<br />

GCs and feedback fibers showed rich spontaneous activity and<br />

diverse excitatory and inhibitory odor responses. On average,<br />

to our stimulus panel (up to 30 odors), we observed 54% purely<br />

excitatory and only 17% purely inhibitory responses in the<br />

GCs, while the two response types were equally common in<br />

the feedback fibers. Interestingly, both GCs and feedback fibers,<br />

that showed spontaneous activity, were inhibited upon odor<br />

presentation, irrespective of stimulus identity. While inhibition<br />

of feedback fibers was sparse and odor specific, GCs showed<br />

both broad, and narrowly tuned inhibitory responses. Further,<br />

a significant fraction (~25%) of GCs exhibited excitatory OFF<br />

responses, independent of stimulus duration. We are currently<br />

investigating how response properties of GCs and feedback<br />

fibers are shaped by odor experience and during rein<strong>for</strong>cement<br />

learning. Additionally, we are employing pharmacological and<br />

optogenetic approaches to modulate the feedback fibers, while<br />

simultaneously monitoring granule cell activity.<br />

#37 SYMPOSIUM:<br />

NEW APPROACHES TO PHYSIOLOGY<br />

AND BEHAVIOR IN AWAKE RODENTS<br />

Odor-guided behaviors in head-restrained and<br />

freely-moving mice<br />

Venkatesh N Murthy 1,2 , Dan Rokni 1,2 , David H. Gire 1,2 ,<br />

Daniel Millman 1,2<br />

1<br />

Harvard University, Molecular & Cellular Biology Cambridge,<br />

MA, USA, 2 Harvard University, Center <strong>for</strong> Brain Science Cambridge,<br />

MA, USA<br />

Olfaction plays a central role in guiding behavior in many<br />

animals, including the common laboratory mammalian models<br />

– rats and mice. There is a renewed excitement about studying<br />

the neural basis of such behaviors, which entails developing<br />

behavioral tasks under controlled conditions where neural<br />

activity can be recorded and manipulated. We have trained mice<br />

to per<strong>for</strong>m odor tasks while their heads are restrained, which<br />

allows stable electrophysiological recordings, high-resolution<br />

optical imaging and optogenetic manipulation. In one such task,<br />

head-restrained mice can be trained to recognize target odorants<br />

embedded in unpredictable and variable background mixtures.<br />

We have also developed strategies to study odor-guided behaviors<br />

in freely moving animals, including spatial navigation. We will<br />

present a detailed analysis of these behaviors in our talk, and<br />

some initial ef<strong>for</strong>ts in recording and imaging neural activity<br />

under these conditions. Acknowledgements: R01DC011291<br />

ORAL ABSTRACTS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

20


#38 SYMPOSIUM:<br />

NEW APPROACHES TO PHYSIOLOGY<br />

AND BEHAVIOR IN AWAKE RODENTS<br />

#39 SYMPOSIUM:<br />

NEW APPROACHES TO PHYSIOLOGY<br />

AND BEHAVIOR IN AWAKE RODENTS<br />

Baseline states and odour evoked responses of mitral and<br />

tufted cells in the awake mouse<br />

Andreas T. Schaefer 1,2,3 , Anja Schmaltz 1,2 , Izumi Fukunaga 1 ,<br />

Mostafa Abdelhamid 1,2 , Mihaly Kollo 1<br />

1<br />

Behavioural Neurophysiology, MPI <strong>for</strong> medical research Heidelberg,<br />

Germany, 2 Dept Anatomy & Cell Biology, University Heidelberg<br />

Heidelberg, Germany, 3 Division of Neurophysiology, MRC-NIMR<br />

London, United Kingdom<br />

Odour stimuli evoke activity patterns in the olfactory bulb,<br />

which are trans<strong>for</strong>med in multiple steps by local microcircuits,<br />

be<strong>for</strong>e arriving to the cortex. Since inhibitory and excitatory<br />

synapses both exhibit short-term plasticity, the baseline activities<br />

and properties of odour-evoked responses strongly determine,<br />

how much individual neurons can influence in<strong>for</strong>mation<br />

processing. There<strong>for</strong>e, gaining an unbiased picture about the<br />

states of different neurons in the behaving animal is essential <strong>for</strong><br />

understanding input trans<strong>for</strong>mation. We have per<strong>for</strong>med blind<br />

whole-cell patch-clamp recordings in awake head-fixed mice to<br />

gain detailed and unbiased measurements of spontaneous and<br />

odour-evoked activity. Mitral and tufted cells (M & TCs) show<br />

great diversity in their baseline resting membrane potentials and<br />

spike rates. Both in awake and anesthetized animals, a large<br />

proportion (>33%) of cells have very low spontaneous firing<br />

rates (


auditory cues anticipating the general availability of taste. Data<br />

from different behavioral paradigms will be used to demonstrate<br />

that both classical and instrumental conditioning can result in<br />

robust anticipatory responses in the GC of restrained as well<br />

as freely moving animals. Experiments investigating the role of<br />

thalamic and limbic inputs in generating cue-responses will be<br />

presented. After a discussion of the system-level underpinnings<br />

<strong>for</strong> the integration of taste-coding and general expectation, I<br />

will show novel evidence that neurons in GC can also learn<br />

to encode specific expectation. Results from neural recordings<br />

in rats per<strong>for</strong>ming a two-cues auditory go/no-go task will<br />

be discussed. I will present data showing that cue responses<br />

are outcome specific and decrease significantly after partial<br />

extinction of cue-taste contingencies. Taste responses will be<br />

analyzed in neurons with different profiles of cue responsiveness<br />

and a relationship between responses to sucrose-anticipating<br />

cues and to sucrose itself will be established. Altogether our<br />

data will further establish the function of GC in integrating<br />

sensory and anticipatory signals and will provide a first analysis<br />

of the system-level mechanisms mediating this function.<br />

Acknowledgements: NIDCD R01-DC010389<br />

#41 PLATFORM PRESENTATIONS: TASTE<br />

Characterization of Testicular Bitter Taste Receptor-Mediated<br />

Signal Transduction<br />

Jiang Xu, Liquan Huang<br />

Monell Chemical Senses Center Philadelphia, PA, USA<br />

Bitter taste receptors were initially isolated from mammalian<br />

taste bud cells in the oral cavity and are believed to function as<br />

a gatekeeper to prevent poisonous substances from ingestion.<br />

In addition to taste buds, however, these receptors have also<br />

been found in some extraoral tissues such as the respiratory<br />

epithelium and gastrointestinal tract. We detected the expression<br />

of these receptors in both human and murine testis, and found<br />

that male germ cells including spermatids and epididymal<br />

sperm can respond to bitter tasting substances by increasing<br />

intracellular calcium concentrations in a dose-dependent manner,<br />

and these calcium responses can be blocked by specific bitter<br />

blockers or the ablation of the G protein gustducin. Further<br />

characterization of these responses with pharmacological agents<br />

indicated that depletion of intracellular calcium stores with<br />

thapsigargin nearly abolished the calcium responses to the bitter<br />

compounds tested whereas the absence of the extracellular<br />

calcium did not affect the response amplitudes. Pretreatment of<br />

the cells with an adenylate cyclase inhibitor, MDL12,330A, also<br />

did not alter the response to caffeine. Thus our data suggested<br />

that testicular bitter taste receptors employ the G protein<br />

gustducin and calcium channels in the endoplasmic reticulum to<br />

mediate calcium responses to bitter tasting compounds. Further<br />

studies are in progress to elucidate the possible roles of these<br />

receptors in the reproductive system. Acknowledgements: This<br />

work was supported by National Institutes of Health Grant R01<br />

DC007487 to L.H., by NIH-NIDCD Core Grant P30 DC011735<br />

to R. Margolskee in support of Monell Core Facilities, and by<br />

National Science Foundation Equipment Grant DBI-0216310 to<br />

N. Rawson in support of Monell’s Confocal Microscopy.<br />

#42 PLATFORM PRESENTATIONS: TASTE<br />

BDNF Maintains Adult Taste Innervation and Is Required<br />

For Taste Nerve Regeneration After Injury<br />

Lingbin Meng, Robin Krimm<br />

University of louisville,ASNB department Louisville, KY, USA<br />

Brain derived neurotropic factor (BDNF) is required <strong>for</strong><br />

the gustatory neuron survival and target innervation during<br />

development. However, whether BDNF has any function in<br />

the adult gustatory system or influences degeneration and<br />

regeneration after nerve injury is unclear. To address these issues,<br />

we inducibly removed BDNF in adulthood. In experimental<br />

animals, Bdnf expression decreased to 4% of control mice in<br />

the lingual epithelium and geniculate ganglion (p


ecome resistant to reward devaluation, i.e. develop habit-like<br />

intake patterns. Adult mice were given 1h ad libitum access<br />

to sucralose paired to intra-gastric infusions of either glucose<br />

or sucralose. After a 14-days conditioning period, animals<br />

were tested after reward devaluation produced by a glucose<br />

intra-gastric preload. Mice challenged with glucose, but not<br />

with sucralose, maintained their daily intake after the preload,<br />

suggesting that sugar but not artificial sweeteners favor the<br />

<strong>for</strong>mation of habits. Accordingly, intra-gastric glucose but not<br />

sucralose produced robust dopamine release in dorsolateral<br />

striatum, suggesting that sugar calories but not sweetness alone<br />

control dopamine efflux in this circuit. Current experiments<br />

involve evaluating whether stimulating dopamine signaling in<br />

dorsolateral striatum is sufficient <strong>for</strong> the <strong>for</strong>mation of artificial<br />

sweetener habitual intake. Our data indicate that the dorsolateral<br />

striatum is critical <strong>for</strong> the inflexible nature associated with sugar<br />

intake. Acknowledgements: NIDCD grant DC009997<br />

#44 PLATFORM PRESENTATIONS: TASTE<br />

CALHM1 ion channel mediates purinergic neurotransmission<br />

of sweet, bitter and umami tastes<br />

J. Kevin Foskett 1 , Akiyuki Taruno 1 , Zhongming Ma 1 , Ichiro<br />

Matsumoto 2 , Michael G. Tordoff 2 , Philippe Marambaud 3<br />

1<br />

Dept Physiology, University of Pennsylvania Philadelphia, PA, USA,<br />

2<br />

Monell Chemical <strong>Sciences</strong> Center Philadelphia, PA, USA, 3 The<br />

Feinstein Institute <strong>for</strong> Medical Research Manhasset, NY, USA<br />

Recognition of sweet, bitter and umami tastes requires the nonvesicular<br />

release from taste bud cells of adenosine 5’-triphosphate<br />

(ATP), which acts as a neurotransmitter to activate afferent<br />

neural gustatory pathways. However, how ATP is released to<br />

fulfill this function is not fully understood. Here we show that<br />

calcium homeostasis modulator 1 (CALHM1), a voltage-gated<br />

ion channel, is indispensable <strong>for</strong> taste stimuli-evoked ATP release<br />

from sweet-, bitter- and umami-sensing taste bud cells. Calhm1<br />

knockout mice have severely impaired perceptions of sweet,<br />

bitter and umami compounds, whereas sour and salty taste<br />

recognition remains mostly normal. Calhm1 deficiency affects<br />

taste perception without interfering with taste cell development<br />

or integrity. CALHM1 is expressed specifically in sweet/bitter/<br />

umami-sensing type II taste bud cells. Its heterologous expression<br />

induces a novel ATP permeability that releases ATP from cells<br />

in response to manipulations that activate the CALHM1 ion<br />

channel. Knockout of Calhm1 strongly reduces voltage-gated<br />

currents in type II cells and taste-evoked ATP release from taste<br />

buds without affecting the excitability of taste cells to taste<br />

stimuli. Thus, CALHM1 is a voltage-gated ATP release channel<br />

required <strong>for</strong> sweet, bitter and umami taste perception.<br />

#45 PLATFORM PRESENTATIONS: TASTE<br />

Sonic Hedgehog Drives Nerve-Independent Formation of<br />

Adult Taste Buds<br />

David Castillo 1,2 , Kerstin Seidel 3 , Ernesto Salcedo 1 , Christina Ahn 4 ,<br />

Frederic J. de Sauvage 4 , Ophir Klein 3,5,6 , Linda Barlow 1,2<br />

1<br />

Dept of Cell and Developmental Biology and the Rocky Mountain<br />

Taste and Smell Center, University of Colorado School of Medicine<br />

Aurora, CO, USA, 2 Graduate Program in Cell Biology, Stem Cells<br />

and Development, University of Colorado School of Medicine Aurora,<br />

CO, USA, 3 Program in Craniofacial and Mesenchymal Biology and<br />

Department of Orofacial <strong>Sciences</strong>, University of Cali<strong>for</strong>nia, San<br />

Francisco San Francisco, CA, USA, 4 Dept. of Molecular Biology,<br />

Genentech Inc South San Francisco, CA, USA, 5 Department of<br />

Pediatrics, University of Cali<strong>for</strong>nia, San Francisco San Francisco, CA,<br />

USA, 6 Institute <strong>for</strong> Human Genetics, University of Cali<strong>for</strong>nia San<br />

Francisco San Francisco, CA, USA<br />

Sonic hedgehog (Shh) is a key regulator of cell proliferation<br />

and differentiation. In the developing tongue, Shh patterns<br />

embryonic taste buds, but its role in maintenance of adult<br />

taste buds is unknown. We used an inducible, Cre-lox system,<br />

where K14CreER, when induced by tamoxifen, drives mosaic<br />

expression of Shh in taste and lingual basal keratinocytes.<br />

Unexpectedly, ectopic Shh expression induced ectopic taste<br />

bud <strong>for</strong>mation in the lingual epithelium, a condition never<br />

encountered in wild type mice. Tongues examined <strong>for</strong> a general<br />

marker of taste cells, cytokeratin (K)8 at 1-4 weeks postinduction,<br />

had numerous K8+ cell clusters found external to<br />

fungi<strong>for</strong>m papillae. Further, using markers of 3 differentiated<br />

taste cell types, we found type I, II and III taste cells in these<br />

ectopic cell clusters, revealing that these clusters were indeed<br />

taste buds. As taste bud differentiation and maintenance are<br />

nerve-dependent, we assessed if <strong>for</strong>mation of ectopic buds was<br />

likewise nerve-dependent. Using P2X2, a marker of gustatory<br />

innervation, abundant neurites were present in fungi<strong>for</strong>m buds,<br />

but were lacking from ectopic buds. However, using PGP9.5, a<br />

general marker of lingual nerve fibers, neurites were found in<br />

the vicinity of ectopic buds, although many ectopic buds were<br />

devoid of innervation. There<strong>for</strong>e, we developed an unbiased,<br />

automated MATLAB script to quantify PGP9.5+ neurites in<br />

and around K8+ taste buds. We found no relationship between<br />

innervation and the presence of ectopic taste buds, controverting<br />

the general consensus that innervation is required <strong>for</strong> adult taste<br />

bud differentiation. Moreover, we demonstrate that Shh supplied<br />

via local epithelial cells is sufficient to drive differentiation of<br />

the entire taste bud cell complement in ectopic locations in a<br />

nerve-independent manner. Acknowledgements: NIH/NIDCD<br />

DC008373 and DC012383 to LAB, and DC004657 to<br />

D. Restrepo<br />

ORAL ABSTRACTS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

23


#46 PLATFORM PRESENTATIONS: TASTE<br />

Phenotypic Characterization of a Caudal to Rostral<br />

Intrasolitary Pathway<br />

Joseph M Breza, Zhixiong Chen, Joseph B Travers, Susan P Travers<br />

The Ohio State University Columbus, OH, USA<br />

Interactions between gustatory and visceral signals modulate<br />

ingestive behavior. The current study explored such interactions<br />

in the nucleus of the solitary tract (NST) where gustatory and<br />

visceral primary afferents terminate in largely separate rostral<br />

(rNST) and caudal (cNST) regions, respectively. Previous<br />

anterograde tracing suggests an intrasolitary projection from<br />

cNST to rNST (Karimnamazi et al.,’98) but the phenotypes<br />

of these connections have not been characterized. We made<br />

electrophysiologically–guided iontophoretic injections of the<br />

retrograde tracer, Fluoro-Gold (FG), into gustatory NST in wildtype<br />

and transgenic mice expressing EGFP under the control<br />

of the GAD67 promoter. Sections were immunostained <strong>for</strong> FG<br />

and tyrosine hydroxylase (TH) to identify catecholaminergic<br />

neurons or PHOX2b, a transcription factor that colocalizes with<br />

a subset of glutamatergic cells. Retrogradely–labeled cNST<br />

neurons extended caudal to obex. At a mid–postremal level,<br />

where several neuron types, including A2 catecholaminergic cells<br />

important in satiety are prominent, the intrasolitary pathway<br />

contained both PHOX2b/excitatory (~65+2.7%) and GAD67/<br />

inhibitory (27+2.5%) projections, with PHOX2b neurons<br />

located preferentially in the medial and intermediate NST and<br />

the GAD67 neurons distributed more laterally, including the<br />

ventral lateral NST. A smaller population (7+1.2%), presumably<br />

a subset of the PHOX2b cells, was double-labeled <strong>for</strong> FG and<br />

TH. Parallel experiments in rats showed a similar distribution of<br />

FG neurons labeled with dopamine beta hydroxylase implying<br />

that most TH–labeled neurons in mice are noradrenergic.<br />

In addition, in vitro recordings from rNST demonstrate that<br />

norepinephrine has a suppressive effect on solitary tract–evoked<br />

responses, suggesting a functional role <strong>for</strong> the A2 projection.<br />

Acknowledgements: Supported by DC00416 & T32DE014320<br />

#47 PLATFORM PRESENTATIONS: TASTE<br />

Longitudinal analysis of 40% calorie restriction on rat taste<br />

bud morphology and expression of sweet taste modulators<br />

Huan Cai, Wei-na Cong, Rui Wang, Caitlin Daimon, Patrick Chirdon,<br />

Rafael deCabo, Stuart Maudsley, Bronwen Martin<br />

National Institute on Aging Baltimore, MD, USA<br />

Taste perception is strongly associated with body weight and<br />

metabolic state. Caloric restriction (CR) is a well-characterized<br />

intervention that reduces body weight and improves metabolic<br />

function. In this study, we per<strong>for</strong>med a longitudinal analysis to<br />

determine the effects of 40% CR on rat taste bud morphology<br />

and expression of sweet taste modulators. Immunohistochemical<br />

analyses of the effects of 40% CR on taste buds were made with<br />

5-, 17- and 30-month old male Fisher 344 rats. No significant<br />

effects of 40% CR on taste bud size and number of taste cells<br />

per taste bud were noted. However, 30-month old rats (both<br />

ad libitum (AL) and calorie restriction (CR) groups) possessed<br />

smaller taste bud size and fewer taste cells per bud than<br />

5- (significant) or 17- (non-significant) month old CR or AL<br />

rats. There was no significant effect of 40% CR or aging on<br />

Type 1 (NTPDase 2), Type 2 (PLC-beta 2), or Type 3 (NCAM)<br />

taste cells marker expression. In contrast, both 40% CR and<br />

AL 30-month old rats demonstrated significantly lower Type 4<br />

(Shh) taste cell marker expression. We found that a-gustducin<br />

expression was significantly higher in 5-month old 40% CR rats<br />

compared to AL, with similar trends <strong>for</strong> T1r3 and glucagonlike<br />

peptide 1 (GLP-1) expression. However, T1r3, GLP-1 and<br />

a-gustducin expression were decreased in 30-month old 40%<br />

CR rats compared to age-matched AL rats. Leptin receptor<br />

expression was significantly higher in 17- and 30-month old 40%<br />

CR rats, compared to age-matched AL rats. Our findings suggest<br />

that short- and long-term CR elicit differential responses on rat<br />

taste bud morphology and sweet taste modulator expression.<br />

This is likely due to long- and short-term calorie intake<br />

and metabolic homeostatic adaptations to the CR regimen.<br />

Acknowledgements: This work was supported entirely by the<br />

Intramural Research Program of the National Institute on<br />

Aging, National Institutes of Health.<br />

#48 SYMPOSIUM:<br />

EXPERIENCE DRIVEN PLASTICITY<br />

OF THE OLFACTORY SYSTEM<br />

Experience Driven Plasticity of the Olfactory System<br />

Xavier Grosmaitre<br />

CSGA, UMR 6265 CNRS - 1324 INRA - Université de Bourgogne<br />

Dijon, France<br />

The olfactory system has been thoroughly investigated during<br />

several decades. A number of its functions have been well<br />

described such as i) the olfactory transduction pathways; ii) the<br />

central odor processing and cerebral and neural networks; iii)<br />

the mechanisms of neurogenesis and synaptic plasticity. The<br />

olfactory system reveals itself as a flexible sensory processing<br />

system. While the implication of experience and internal state<br />

on olfactory detection and processing has been explored, some<br />

points remain unclear: whether experience can modulate specific<br />

populations of olfactory sensory neurons (OSNs) and olfactory<br />

bulb (OB) input and output maps. New tools were recently<br />

developed to analyze the plasticity of the olfactory system such<br />

as genetic labeling of specific population of OSNs (i.e. expressing<br />

specific OR), molecular biology tools (qRT-PCR), genetic<br />

labeling <strong>for</strong> synaptic transmission imaging and in vivo imaging.<br />

This symposium focuses on the dynamic nature of olfactory<br />

processing: the effects of experience on how odors are processed<br />

from olfactory sensory neurons to higher-order structures and<br />

how recent techniques are used to investigate these questions in<br />

different structures and models. In mice, we will discuss: i) the<br />

effects of olfactory experience on specific OSNs populations and<br />

synaptic transmission; ii) the effects of emotional experience<br />

on OSNs function; ii) the importance of the balance between<br />

activity, regeneration and remodeling <strong>for</strong> OB plasticity; iii) the<br />

use of long-term imaging of odor representations in awake mice<br />

to explore olfactory plasticity. Finally, plasticity induced by<br />

ORAL ABSTRACTS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

24


olfactory experience on behavior and brain in an Insect model<br />

will be presented. We aim to present the state of the art about the<br />

olfactory system as a flexible, adaptable and plastic system.<br />

#50 SYMPOSIUM:<br />

EXPERIENCE DRIVEN PLASTICITY<br />

OF THE OLFACTORY SYSTEM<br />

#49 SYMPOSIUM:<br />

EXPERIENCE DRIVEN PLASTICITY<br />

OF THE OLFACTORY SYSTEM<br />

Postnatal Odorant Exposure Induces Peripheral<br />

Olfactory Plasticity<br />

Xavier Grosmaitre<br />

CSGA, UMR 6265 CNRS - 1324 INRA - Université de Bourgogne<br />

Dijon, France<br />

Olfactory sensory neurons (OSNs) <strong>for</strong>m an interface between<br />

the environment and the brain, converting chemical in<strong>for</strong>mation<br />

(odorants) into electrical signals and sending these signals<br />

to the brain. Little is known about the consequences of long<br />

term odorant exposure on OSNs. Our goal is to understand<br />

the anatomical, molecular and physiological effects of odorant<br />

exposure at the cellular level and more precisely in the context<br />

of an early postnatal olfactory exposure. We focus our work on<br />

specific populations of OSNs expressing particular ORs using<br />

gene-targeted mice. MOR23-GFP mice were exposed daily to<br />

Lyral and anatomical, molecular and physiological properties<br />

of these neurons were analyzed. The density of MOR23<br />

neurons decreased after odorant exposure while the level of<br />

mRNA <strong>for</strong> the receptor remained stable at the entire mucosa<br />

level. To investigate molecular changes within individual OSNs,<br />

mRNA levels <strong>for</strong> olfactory signaling pathway components<br />

were quantitatively analyzed using qPCR on GFP-containing<br />

neurons (7 per mouse). The levels of mRNAs <strong>for</strong> CNGA2,<br />

PDE1C and MOR23 olfactory receptor were higher in exposed<br />

OSNs compared to control. Using patch-clamp recordings on<br />

the dendritic knobs of MOR23 neurons in an intact preparation<br />

we observed that exposed OSNs displayed a lower detection<br />

threshold compared to control OSNs while the dynamic range<br />

of the dose-response was broader. Responses of exposed neurons<br />

were also faster and shorter than the responses of control<br />

neurons. Postnatal odorant exposure induces molecular and<br />

physiological plasticity in individual MOR23 neurons. Taken<br />

together, our data suggest that the olfactory epithelium presents<br />

deep anatomical, molecular and functional changes when<br />

chronically exposed to odorant molecules in early stage of life.<br />

Acknowledgements: Funding was provided by CNRS (ATIP<br />

grant), by Conseil Régional de Bourgogne (FABER and PARI<br />

grants), and by Université de Bourgogne (BQR program).<br />

Altered olfactory sensory neuron physiology following odor<br />

exposure or olfactory fear conditioning in vivo<br />

John P. McGann, Michelle C. Rosenthal, Marley D. Kass,<br />

Andrew H. Moberly<br />

Rutgers University / Psychology Department Piscataway, NJ, USA<br />

Experience-dependent plasticity is increasingly understood to<br />

occur throughout adulthood in mammalian sensory systems.<br />

This talk will present data from experiments that used optical<br />

imaging to longitudinally assess the effects of odorant exposure<br />

and emotional learning on the physiology of olfactory<br />

sensory neurons (OSNs) in vivo in individual adult mice. In 1<br />

experiment, mice expressing the fluorescent exocytosis indicator<br />

synaptopHluorin in mature OSNs underwent a baseline imaging<br />

session in which a chronic cranial window was implanted in<br />

the skull overlying the olfactory bulbs and the OSN synaptic<br />

output into olfactory bulb glomeruli was visualized during the<br />

presentation of a panel of 4 odorants (2 esters, 1 aldehyde, and<br />

1 ketone). Mice then spent a week in either an odorant-exposure<br />

chamber in which 1 of the esters was presented repeatedly with<br />

a 4 hour duty cycle or in a control chamber. After exposure,<br />

the exposed ester and the control ester both evoked less OSN<br />

synaptic output into fewer glomeruli than prior to exposure,<br />

while aldehyde- and ketone-evoked responses were unchanged.<br />

In a separate experiment, mice underwent a similar baseline<br />

imaging session but were then differentially fear conditioned<br />

to associate 1 ester (the CS+) with shock while the other ester<br />

(the CS-) was presented without shock. In a post-conditioning<br />

imaging session the glomeruli that received OSN input driven<br />

by the CS+ received larger synaptic inputs from OSNs, while<br />

responses to the CS- and non-presented control odorants were<br />

unchanged. Control mice that experienced only odors or only<br />

shocks between imaging sessions exhibited no changes in the<br />

response to any odorant. These data provide surprising evidence<br />

that environmental changes and emotional learning can change<br />

how OSNs respond to olfactory stimuli. Acknowledgements:<br />

This work was supported by the National Institute on Deafness<br />

and Other Communication Disorders (R00 DC009442 to JPM).<br />

ORAL ABSTRACTS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

25


#51 SYMPOSIUM:<br />

EXPERIENCE DRIVEN PLASTICITY<br />

OF THE OLFACTORY SYSTEM<br />

#53 SYMPOSIUM:<br />

EXPERIENCE DRIVEN PLASTICITY<br />

OF THE OLFACTORY SYSTEM<br />

Understanding Plasticity in the Olfactory Intrabulbar Map<br />

Leonardo Belluscio<br />

National Institutes of Health / NINDS Bethesda, MD, USA<br />

In the mammalian olfactory system sensory neurons project<br />

their axons to the surface in the olfactory bulb generating a<br />

pair of glomerular maps that reflect odorant receptor identity.<br />

These maps are further connected through a set of reciprocal<br />

intrabulbar projections that are mediated by tufted cells that<br />

specifically link iso-functional odor columns to produce a second<br />

order map called the intrabulbar map. We have shown that<br />

intrabulbar projections are established postnatally and undergo<br />

continuously refinement through an activity dependent process<br />

that has no critical period. Here we present that both loss of<br />

olfactory sensory input and broad odorant stimulation are<br />

capable of disrupting the intrabulbar map specificity, while<br />

re-introduction of normal activity restores the map to proper<br />

order. We also reveal that the regenerating interneurons<br />

are central to intrabulbar circuit plasticity and that proper<br />

connectivity depends specifically upon new neurons from<br />

the rostral migratory stream. Together these data illustrate<br />

that olfactory bulb plasticity is a balance between activity,<br />

regeneration and remodeling. Acknowledgements: National<br />

Institute of Neurological Disorders and Stroke, Intramural<br />

Research Program.<br />

#52 SYMPOSIUM:<br />

EXPERIENCE DRIVEN PLASTICITY<br />

OF THE OLFACTORY SYSTEM<br />

Long-term imaging of odor representations in awake mice<br />

Takaki Komiyama<br />

University of Cali<strong>for</strong>nia, San Diego, CNCB, La Jolla CA 92093 USA<br />

How are sensory representations in the brain influenced by the<br />

state of an animal? Here we use chronic two-photon calcium<br />

imaging to explore how wakefulness and experience shape odor<br />

representations in the mouse olfactory bulb. Comparing the<br />

awake and anesthetized state, we show that wakefulness greatly<br />

enhances the activity of inhibitory granule cells and makes<br />

principal mitral cell odor responses more sparse and temporally<br />

dynamic. In awake mice, brief repeated odor experience leads<br />

to a gradual and long-lasting (months) weakening of mitral cell<br />

odor representations. This mitral cell plasticity is odor specific,<br />

recovers gradually over months, and can be repeated with<br />

different odors. Furthermore, the expression of this experiencedependent<br />

plasticity is prevented by anesthesia. Together, our<br />

results demonstrate the dynamic nature of mitral cell odor<br />

representations in awake animals, which is constantly shaped by<br />

recent odor experience.<br />

Olfactory Experience Shapes Insect Olfactory Centres<br />

Jean-Marc Devaud<br />

Research Center on Animal Cognition, Université Paul Sabatier<br />

Toulouse, France<br />

Insects provide excellent models to study how neural<br />

networks dedicated to olfactory processing are <strong>for</strong>med during<br />

development, and how they work in the adult. However, their<br />

organisation is not fixed once development is achieved.<br />

On the contrary, as in vertebrates, the connectivity and functional<br />

organisation of insect olfactory systems are not fixed: they<br />

exhibit clear plastic properties, as shown in various species over<br />

the recent years. In our work, we have been focusing on the<br />

plastic changes affecting the anatomy of the olfactory centres as<br />

a consequence of olfactory experience, be it the mere exposure<br />

to environmental odorants or associative learning and memory.<br />

In particular, we have been looking <strong>for</strong> structural rearrangements<br />

in two main olfactory centres known <strong>for</strong> their role in olfactory<br />

learning and memory in the insect brain: the antennal lobes and<br />

the mushroom bodies. The modular organization of these two<br />

neuropils allows quantifying the changes affecting their structure<br />

in the brains of animals submitted to different treatments.<br />

By doing so, and by focusing mostly on the honeybee (Apis<br />

mellifera) as a model species, we have been able to show that<br />

the <strong>for</strong>mation of long-term memories of previous olfactory<br />

experience is associated with structural modifications in insect<br />

olfactory networks. Interestingly, such modifications vary with<br />

the nature of the experience undergone by the animal, and<br />

may be considered as supports of olfactory memories. Thus,<br />

they are likely to contribute to the acquisition and retention of<br />

behavioural responses adapted to changing environments.<br />

#53.5 CLINICAL LUNCHEON:<br />

TASTE RECEPTORS IN GUT AND PANCREAS<br />

REGULATE ENDOCRINE FUNCTION<br />

Robert F. Margolskee<br />

Monell Chemical Senses Center, Philadelphia, PA, USA<br />

Many of the receptors and downstream signalling proteins<br />

involved in taste detection and transduction are expressed also<br />

in intestinal and pancreatic endocrine cells where they underlie<br />

certain chemosensory responses. Intestinal endocrine cells<br />

express T1r taste receptors, the taste G-protein gustducin, and<br />

several other taste transduction elements. So too do pancreatic<br />

islet endocrine cells. Knockout mice lacking alpha-gustducin<br />

or the sweet taste receptor subunit T1r3 have deficiencies in<br />

intestinal secretion of glucagon-like peptide-1 (GLP-1) and in<br />

the regulation of plasma levels of insulin and glucose. Glucosedependent<br />

insulin release from mouse pancreatic islets ex vivo<br />

is stimulated by sucralose and other sweeteners. Islets from T1r3<br />

knockout mice release insulin normally in response to glucose,<br />

but show no enhanced release of insulin in response to noncaloric<br />

sweeteners. Thus, there appear to be two mechanisms<br />

ORAL ABSTRACTS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

26


<strong>for</strong> regulating insulin release from pancreas, one dependent on<br />

glucose, glucose transporters and glucokinase, and another <strong>for</strong><br />

sweeteners that involves T1r3 and other taste proteins. Only<br />

limited studies with humans have been done in this area, but it<br />

seems likely that “taste” signalling proteins in human gut and<br />

pancreas also contribute to chemosensory responses in gut and<br />

pancreas to regulate glucose homeostasis.<br />

#54 SYMPOSIUM:<br />

THE NEW ‘FACES’ OF CHEMOSENSATION —<br />

UTILIZING CHEMOSENSORY SIGNALING<br />

PATHWAYS OUTSIDE THE CANONICAL<br />

CHEMOSENSORY ORGANS<br />

Symposium: The New ‘Faces’ of Chemosensation –<br />

Utilizing Chemosensory Signaling Pathways Outside the<br />

Canonical Chemosensory Organs<br />

Yehuda Ben-Shahar 1,2<br />

1<br />

Washington University St. Louis, MO, USA, 2 Washington University<br />

School of Medicine/Pulmonary & Critical Care Medicine St. Louis,<br />

MO, USA<br />

In recent years, several exciting studies indicated that canonical<br />

mammalian chemosensory signaling pathways are also likely<br />

to function outside the canonical chemosensory organs (taste<br />

and olfaction). These findings significantly expand our field of<br />

view in terms of what constitutes a chemosensory organ, and<br />

suggest that cell-autonomous chemoreception, independent<br />

of the nervous system, is likely playing an important role in<br />

health and disease, which includes digestive, respiratory, and<br />

reproductive functions. This symposium will present the current<br />

state of knowledge about the possible extra-sensory functions of<br />

‘taste’ and ‘olfaction’ signaling molecules in non-sensory tissues<br />

and cells.<br />

and primary airway cultures to identified candidate olfactory<br />

sensory cells in human airways, which can respond to volatiles<br />

in culture. Using several well-established markers <strong>for</strong> various<br />

pulmonary epithelial cell types we identified the olfactory cells<br />

as pulmonary neuroendocrine cells (PNEC). In humans, PNECs<br />

are morphologically distinct cells of unknown function. We<br />

found that human PNECs express members of the olfactory<br />

receptor family and are anatomically positioned to respond to<br />

inhaled volatile chemicals. Further, human olfactory PNECs<br />

showed high levels of vesicular 5-HT and the peptide hormone<br />

CGRP, further establishing the pulmonary neuroendocrine<br />

system as olfactory sensory sentinels. Apical exposure of<br />

primary human airway cultures to volatile chemicals led to the<br />

release of the neuroendocrine content of PNECs, indicating<br />

that adult human PNECs could act as olfactory sensory cells.<br />

Since pulmonary tissues express diverse serotonin and peptide<br />

receptors, these data indicate that human airway epithelia<br />

evolved a specialized group of cells that can act autonomously in<br />

response to volatile chemical insults. These cells may represent<br />

the missing cellular and physiological links between the exposure<br />

to environmental volatiles and airway hypersensitivity observed<br />

in some pulmonary diseases. Acknowledgements: NIH NIDCD<br />

R03DC010244 and the Children’s Discovery Institute (St. Louis)<br />

#56 SYMPOSIUM:<br />

THE NEW ‘FACES’ OF CHEMOSENSATION —<br />

UTILIZING CHEMOSENSORY SIGNALING<br />

PATHWAYS OUTSIDE THE CANONICAL<br />

CHEMOSENSORY ORGANS<br />

Bitter Taste Receptors on Airway Smooth Muscle:<br />

a target <strong>for</strong> novel bronchodilators<br />

Stephen Liggett<br />

University of South Florida Morsani College of Medicine. Tampa, FL<br />

ORAL ABSTRACTS<br />

#55 SYMPOSIUM:<br />

THE NEW ‘FACES’ OF CHEMOSENSATION —<br />

UTILIZING CHEMOSENSORY SIGNALING<br />

PATHWAYS OUTSIDE THE CANONICAL<br />

CHEMOSENSORY ORGANS<br />

Human Pulmonary Neuroendocrine Cells are<br />

Olfactory Sentinels<br />

Yehuda Ben-Shahar 1,2<br />

1<br />

Washington University St. Louis, MO, USA, 2 Washington University<br />

School of Medicine/Pulmonary & Critical Care Medicine St. Louis,<br />

MO, USA<br />

Previous work indicated that mammalian pulmonary ciliated<br />

epithelial cells act as chemosensory cells to non-volatile<br />

stimuli. Whether human airways can also detect volatiles<br />

is not known. However, human pulmonary diseases such<br />

as asthma have been linked to increased sensitivity of the<br />

airways to various volatile insults. We have identified several<br />

expressed canonical olfactory receptors in human primary<br />

airway epithelia. Immunohistochemistry on lung sections<br />

There is an unmet need <strong>for</strong> treatments of the airway constriction<br />

that occurs in asthma. Currently, beta-agonists are the only class<br />

of direct bronchodilators that are in use. We unexpectedly found<br />

T2Rs (particularly subtypes 10, 14, 31) expressed on human<br />

airway smooth muscle (ASM). Activation by T2R agonists<br />

causes significant and reversible ASM relaxation in human and<br />

mouse airways, isolated human ASM cells studied by magnetic<br />

twisting cytometry and Fourier trans<strong>for</strong>m traction maps, and in<br />

mice in vivo. Signaling was sensitive to inhibitors of betagamma,<br />

PLC, and the IP3 receptor, and was associated with membrane<br />

hyperpolarization and an increase in SR-released intracellular<br />

[Ca2+] within a sequestered pool restricted to the cell surface.<br />

One channel that appears to be involved is BKca, but there may<br />

be others. In the ovalbumin-sensitized mouse model of asthma,<br />

the inhaled T2R agonist quinine was much more efficacious<br />

than the inhaled beta agonist albuterol in reversing airflow<br />

obstruction. Given the large number of known bitter tastants,<br />

there is an opportunity to discover non-toxic T2R agonists <strong>for</strong><br />

inhalation <strong>for</strong> the treatment of obstructive lung disease.<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

27


#57 SYMPOSIUM:<br />

THE NEW ‘FACES’ OF CHEMOSENSATION —<br />

UTILIZING CHEMOSENSORY SIGNALING<br />

PATHWAYS OUTSIDE THE CANONICAL<br />

CHEMOSENSORY ORGANS<br />

Genetics of the bitter taste receptor T2R38 underlie<br />

susceptibility to upper respiratory infection<br />

Noam A. Cohen 1,2 , Robert J. Lee 1 , Danielle R. Reed 3 , Peihua Jiang 3 ,<br />

Gary K. Beauchamp 3<br />

1<br />

University of Pennsylvania/Otorhinolaryngology - Head and Neck<br />

Surgery Philadelphia, PA, USA, 2 Philadelphia VA Medical Center/<br />

Surgery Philadelphia, PA, USA, 3 Monell Chemical Senses Center<br />

Philadelphia, PA, USA<br />

Innate and adaptive defense mechanisms protect the respiratory<br />

system from attack by microbes. Here, I describe our recent<br />

studies that demonstrate that the bitter taste receptor T2R38<br />

is expressed in the human upper respiratory epithelium and is<br />

activated by acyl-homoserine lactone quorum sensing molecules<br />

secreted by Pseudomonas aeruginosa and other gram-negative<br />

bacteria. T2R38 regulates human upper airway innate defenses<br />

through nitric oxide production, resulting in stimulation of<br />

mucociliary clearance and direct antibacterial effects. Moreover,<br />

common polymorphisms of the TAS2R38 gene are linked to<br />

significant differences in the ability of upper respiratory cells<br />

to clear and kill bacteria. Lastly, TAS2R38 genotype correlates<br />

with human sinonasal gram-negative bacterial infection. These<br />

data suggest that T2R38 is an upper airway sentinel in innate<br />

defense, and that genetic variation that contributes to human<br />

individual differences in ability to taste phenylcarbamide (PTC)<br />

and related molecules also contributes to individual differences<br />

in susceptibility to respiratory infection. Acknowledgements:<br />

P30DC011735 R01DC004698 P50DC000214 R01DC010842<br />

#58 SYMPOSIUM:<br />

THE NEW ‘FACES’ OF CHEMOSENSATION —<br />

UTILIZING CHEMOSENSORY SIGNALING<br />

PATHWAYS OUTSIDE THE CANONICAL<br />

CHEMOSENSORY ORGANS<br />

Detection Of Irritants And Bacterial Metabolites Via<br />

The Taste Transduction Cascade In Solitary Chemosensory<br />

Cells Of The Nasal Cavity<br />

Thomas Finger 1,3 , Sue Kinnamon 2,3 , Vijay Ramakrishnan 2 ,<br />

Marco Tizzano 1,3<br />

1<br />

Dept. Cell & Devel. Biol., Univ Colo. Med. Sch. Aurora, CO, USA,<br />

2<br />

Dept. Otolaryngology, Univ Colo. Med. Sch. Aurora, CO, USA, 3 Rocky<br />

Mountain Taste & Smell Center Aurora, CO, USA<br />

Airways are continually assaulted by harmful compounds carried<br />

on the incoming airstream and by the potentially pathogenic<br />

bacterial populations. We have shown that the nasal epithelium<br />

of rodents houses a population of trigeminally-innervated<br />

solitary chemosensory cells (SCCs) that express T2R taste<br />

receptors along with their downstream signaling components<br />

crucial <strong>for</strong> detection and response to these deleterious agents.<br />

SCCs are distributed across much of the nasal respiratory<br />

epithelium in rodents, but are especially concentrated along<br />

curved surfaces facing major airway conduits. In the upper<br />

respiratory passageways, the SCCs express nearly all T2Rs tested<br />

as well as some T1R family members. Downstream signaling<br />

components also are expressed by SCCs including: gustducin,<br />

PLCß2 and TrpM5. Functionally, SCCs respond rapidly to bitter<br />

ligands (e.g. denatonium 10mM) as well as bacterial signaling<br />

molecules including the acylhomoserine lactones produced as<br />

quorum sensing molecules by Pseudomonas species. Activation<br />

of the SCCs by these compounds evokes neurally-mediated<br />

protective reflexes, involving both apnea and local neurogenic<br />

inflammation. These responses are absent in both gustducin<br />

and TrpM5 knockout mice implicating the taste transduction<br />

cascade as a crucial component of responses to these substances.<br />

Moreover, the local inflammatory responses were absent after<br />

chemical ablation of the nociceptive fibers of the trigeminal<br />

nerve showing the necessity <strong>for</strong> innervation. More recently we<br />

have tested the possibility that similar SCCs are present in the<br />

human nose and now report (S. Cooper AChemS 2013) the<br />

presence of elongate TrpM5+ microvillous cells within the<br />

sinonasal epithelium in humans. Whether similar SCCs underlie<br />

epithelial defense systems in humans remains to be determined.<br />

Acknowledgements: Supported by NIDCD grants to T. Finger,<br />

SC Kinnamon (RO1 DC009820), M Tizzano (R03 DC012413)<br />

and the Rocky Mountain Taste & Smell Center (P30 DC04657)<br />

#59 SYMPOSIUM:<br />

THE NEW ‘FACES’ OF CHEMOSENSATION —<br />

UTILIZING CHEMOSENSORY SIGNALING<br />

PATHWAYS OUTSIDE THE CANONICAL<br />

CHEMOSENSORY ORGANS<br />

Block of taste genes leads to male sterility<br />

Bedrich Mosinger, Kevin M. Redding, Rockwell M. Parker,<br />

Robert F. Margolskee<br />

Monell Chemical Senses Center Philadelphia, PA, USA<br />

The male infertility rate in the developing world is increasing.<br />

About 7% of men are infertile <strong>for</strong> unknown reasons, while<br />

having normal hormone levels. We found that two gene products<br />

(T1R3 and gustducin) originally found in taste chemosensation<br />

are expressed in haploid spermatids and their absence or<br />

block leads to male sterility. The pathology shows an arrest in<br />

spermatid development with numerous giant and exfoliated<br />

cells in testicular tubules, oligospermia and immotile sperm.<br />

To study this phenomenon we produced a mouse model<br />

expressing a humanized <strong>for</strong>m of T1R3 that can be inhibited<br />

by human-specific inhibitors. Some of these inhibitors are<br />

phenoxy compounds developed into common medications and<br />

agricultural chemicals. A very effective, yet reversible, sterility<br />

is achieved by treatment with the phenoxy compound clofibrate<br />

in this model with humanized T1R3 receptor in the background<br />

of gustducin null allele. Because both T1R3 and gustducin affect<br />

cAMP levels, we hypothesize that their absence impairs function<br />

of CREM (cAMP response modulator protein) a transcriptional<br />

master gene indispensable <strong>for</strong> spermatid development and EPAC<br />

(exchange protein activated by cAMP) a Rap 1 activator required<br />

ORAL ABSTRACTS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

28


<strong>for</strong> cell adhesion. We further hypothesize that even low levels of<br />

chemicals and/or medications acting on T1R3 and gustducin,<br />

possibly in combination with other drugs, can negatively affect<br />

human male fertility. Acknowledgements: R21 DC007399 -<br />

NIH/NIDCD, R01 DC003155 - NIH/NIDCD, P30DC011735<br />

- NIH/NIDCD.<br />

# 60 IFF LECTURE:<br />

BITTER TASTE IN MICE AND MAN<br />

Bitter Taste in Mice and Man<br />

Wolfgang Meyerhof<br />

Department of Molecular Genetics, German Institute of Human<br />

Nutrition, Nuthetal, Germany<br />

Bitterness is elicited by numerous structurally diverse molecules<br />

which usually act as repellents allowing us to avoid intake<br />

of harmful food. However, we adapt to and even prefer the<br />

bitterness of some foods and beverages. To elucidate these<br />

phenomena my laboratory investigates the receptors and cells <strong>for</strong><br />

bitter tasting compounds in man and mice.<br />

ORAL ABSTRACTS<br />

Functional assays with >100 chemicals demonstrate similar<br />

molecular receptive ranges <strong>for</strong> the repertoires of murine and<br />

human Tas2r bitter taste receptors even though the number of<br />

cognate compounds discovered <strong>for</strong> the human TAS2Rs was<br />

larger. This is surprising since the mouse genome encodes 30%<br />

more bitter receptors than the human genome does. Notably,<br />

the two species usually detect the same compounds with nonorthologous<br />

Tas2rs. Both species possess generalists, moderately<br />

tuned Tas2rs and specialists according to their number of<br />

cognate bitter compounds. However, a larger fraction of<br />

specialists occurs in mice, proposing that the luxury of having<br />

specific receptors <strong>for</strong> solitary bitter compounds in a species is<br />

supported by a greater number of TAS2R genes.<br />

Genetic labeling and in situ hybridization experiments visualized<br />

the populations of bitter-sensing cells in mice and man. They<br />

express the complete repertoires of Tas2r genes but individual<br />

cells express only limited subsets. This is supported by genetic<br />

ablation in mice of the cells <strong>for</strong> Tas2r131 which extinguished<br />

only ~50% of the bitter cell population. Oral administration of<br />

bitter compounds to mice excites specifically 200-400 neurons in<br />

the gustatory part of the nucleus of the solitary tract as indicated<br />

by induction of immediate early gene. These gustatory neurons<br />

respond differently to different oral bitter stimuli. Finally, we<br />

found that mice without Tas2r131 cells avoid several bitter<br />

compounds less than controls, whereas they are indistinguishable<br />

from controls in their avoidance of denatonium benzoate.<br />

Our data demonstrate that bitter sensing cells and their gustatory<br />

target neurons are functionally distinct <strong>for</strong>ming the basis <strong>for</strong><br />

variable behavioral responses to different bitter chemicals which<br />

could be of relevance <strong>for</strong> ingestive behavior.<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

29


Poster Presentations<br />

#P1 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Meteorological Vaticination through Phantosmia:<br />

A Case Study<br />

Gurprit Bains, Salvatore Aiello, Alan Hirsch<br />

Smell & Taste Treatment and Research Foundation Chicago, IL, USA<br />

Introduction: Linkage of weather to chemosensory<br />

hallucinations has hereto<strong>for</strong>e not been reported.<br />

Methods: Case Study: A 64 yo M presents with five years<br />

of intermittent noxious skunk-onion excrement phantosmia,<br />

lasting <strong>for</strong> hours. If severe, he tastes the same odor. It is<br />

exacerbated by coughing and nasal congestion, and alleviated<br />

with sleep, nasal irrigation, alprazolam, occluding nostrils,<br />

assuming Moffitt’s position, snorting salt water, blowing nose,<br />

and holding breath. When eating or sniffing, the actual flavors<br />

replace the phantosmia. Since onset, he noted the intensity and<br />

frequency of the phantosmia <strong>for</strong>ecasted the weather. Two hours<br />

be<strong>for</strong>e a storm, the phantosmia intensifies from a level 0 to a<br />

7-10, which persists through the entire thunderstorm. Twenty<br />

years prior, he noted the ability to <strong>for</strong>ecast the weather, based on<br />

pain in a torn meniscus, which vanished after surgical repair.<br />

Results: Olfactory testing revealed hyposmia: QSIT 1; ETOH<br />

Sniff Test 3cm; BSIT 8; PST 2; Odor Memory Test: 2-10 sec.,<br />

2-30 sec., 0-60 sec.; UPSIT – 27 R, 8 L, Sniffin’ Sticks –<br />

Threshold: L


pressure and blood in cerebrospinal fluid. CONCLUSIONS:<br />

The chemosensory loss of the patient is postulated to be due to<br />

deposition of hemosiderin on the olfactory nerve. In superficial<br />

siderosis, anosmia or hyposmia is common, but olfactory<br />

function testing is rarely undertaken. Earlier case reports of<br />

superficial siderosis have not described detailed chemosensory<br />

tests. This is the first report of superficial siderosis-induced<br />

chemosensory deficit with presentation of a set of objective<br />

chemosensory testings.<br />

#P5 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Tests of Retronasal Smell in Children: Which Flavored Jelly<br />

Bean Works Best?<br />

Noah Hirsch 1 , Richard Bone 2 , Alan R Hirsch 1<br />

1<br />

Smell & Taste Treatment and Research Foundation Chicago, IL, USA,<br />

2<br />

Christ Advocate Medical Center Oak Lawn, IL, USA<br />

#P4 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Postviral Hyposmia with Transient Improvement by<br />

Spontaneous Yawning<br />

Alan R Hirsch, Jason Gruss, Gul Hwang<br />

Smell & Taste Treatment and Research Foundation Chicago, IL, USA<br />

Objective: To describe a patient with hyposmia whose olfactory<br />

ability returns upon yawning. Method: 61yo male, three years<br />

ago noted nasal congestion, followed by loss of smell and taste.<br />

Sinus CT showed opacification of the ethmoidal and an airfluid<br />

level in the right maxillary sinus. After treatment, sinus<br />

CT, MRI, and fiberoptic endoscopy were normal, but smell<br />

ability remained at 5%. He is able to smell flowers and gasoline<br />

if held close. Immediately after yawning he was able to smell<br />

<strong>for</strong> a second. With small yawns 20% returns; 100% with large<br />

yawns. Two years ago transient phantosmias began of rubber or<br />

chemicals. His taste is 1-2%. He can taste Chinese mustard and<br />

red pepper. Results: Chemosensory testing indicated hyposmia<br />

and hypogeusia. Q-SIT 2; Sniff Magnitude with Sniff Magnitude<br />

Ratio of .87; Alcohol Sniff Test 11 cm; Sniffin’ Stick Threshold<br />

L -2.0; After Induced natural yawn > -2.0.<br />

Conclusion: Yawning may improve olfaction by enhancing nasal<br />

airflow. The polite yawning technique should be considered to<br />

be part of the evaluation in those who complain of<br />

chemosensory dysfunction.<br />

Objective: In 13 adults, Mozell et al (1969) found flavor<br />

dependent effects of different foods of retronasal smell on<br />

identification, but did not address intensity. Engen (1982) posited<br />

that the olfactory role in identification differs from intensity<br />

assessments. We looked to determine, in children, the retronasal<br />

component of intensity of flavored jelly beans. Methods: Forty<br />

two, 12 or 13 year olds, were screened with the Brief Smell<br />

Identification test. Twenty five (17 girls, 8 boys) scored 3/3 and<br />

underwent an assessment of intensity, on a 10 point visual analog<br />

scale, with and without nose clips, of 10 different Jelly Belly<br />

jelly beans. The mean differences were determined. Each jelly<br />

bean flavor was provided to subjects prior to testing and subjects<br />

verified they had knowledge of what these flavors should taste<br />

like. Results: Significant differences(p


For patients not involved in litigation, we found significant<br />

relationships between odor ID and ratings (p


normal. Allergy skin test was positive <strong>for</strong> garlic and onion.<br />

Nose plug and counter stimulation with peppermint prevented<br />

the onset of migraine. Conclusion: This is the first report of<br />

migraines triggered by more than one alliaceous compound in<br />

the same individual. Possible mechanisms include odor induced:<br />

emotional change; vasomotor instability; trigeminal induced<br />

neurogenic inflammation; and allergic response. In alliaceous<br />

and odor-induced migraines, a trial of counter stimulation and<br />

nose plugs is warranted.<br />

#P11 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

The WUTC Odor Threshold Test: Evaluating Olfactory<br />

Ability Using Signal Detection Theory<br />

William Tewalt, Irene N Ozbek, McKinney Jessica, Santiago Manuel,<br />

Biderman Michael<br />

University of Tennessee at Chattanooga Chattanooga, TN, USA<br />

#P10 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Dysautonomia and Chemosensory Dysfunction<br />

Noorussabah Shaikh 1 , Alan R. Hirsch 1,2,3,4<br />

1<br />

Smell and Taste Treatment and Research Foundation Chicago, IL,<br />

USA, 2 Mercy Hospital and Medical Center/ Department of Medicine<br />

Chicago, IL, USA, 3 Rush University Medical Center/ Department<br />

of Neurology Chicago, IL, USA, 4 Rush University Medical Center/<br />

Department of Psychiatry Chicago, IL, USA<br />

Introduction: Myriad disorders associated with dysautonomias<br />

also display chemosensory dysfunction. Given this overlap,<br />

assessments of autonomic dysfunction amongst patients of a<br />

specialized chemosensory clinic were per<strong>for</strong>med. Methods: 25<br />

consecutive patients at a chemosensory clinic were approached<br />

to participate in this IRB approved study. 22 consented. Average<br />

age 51 years (range 22 - 69) 8 males and 14 females, with<br />

diagnosis of hyposmia/ anosmia (18), dysosmis (3), phantosmia<br />

(6), palinosmia (4), hypogeusia/ ageusia (15), dysgeusia (9),<br />

phantogeusia (9), burning mouth syndrome (4). All underwent<br />

the Quick Smell Identification Test (QSIT), and the Survey of<br />

Autonomic Symptoms (SAS) <strong>for</strong> symptoms and Total Impact<br />

Score (TIS) <strong>for</strong> severity. Results: Our patients demonstrated<br />

fewer autonomic symptoms than either those with autonomic<br />

disorders or even published controls. There was a significant<br />

difference in SAS score in the group as a whole (1.82) compared<br />

to published control SAS of ≤3.0 (p= 0.003) and of the group<br />

as a whole (4.32) compared to TIS of controls ≤ 7.0 ( p< 0.006).<br />

No significance difference between QSIT score and either<br />

SAS or TIS was observed (p>0.1). Discussion: This lack of<br />

autonomic dysfunction in those chemosensory disorders may<br />

reflect a sampling error. Those with autonomic dysfunction<br />

may be so burdened by their primary illness they did not seek<br />

care <strong>for</strong> their chemosensory problems, if they even recognized<br />

them at all. Alternatively, the SAS and TIS may not be valid in<br />

this patient group, while physiologic autonomic testing might<br />

demonstrate dysfunction. A larger sample size may have revealed<br />

dysautonomia. Conclusion: Lack of autonomic symptoms<br />

were seen in those with chemosensory dysfunction. Further<br />

investigation of such a connection is warranted.<br />

A new odor detection threshold test (WUTC) was developed,<br />

using signal detection theory, to address the need <strong>for</strong> establishing<br />

a measure of consistency of subjects’ responses, and the need<br />

<strong>for</strong> introducing different odorants to address different clinical/<br />

research questions. The method of administration used is<br />

an improvement over previous methods of testing because<br />

the likelihood of desensitization to a single odor over time is<br />

minimized. Odorants were chosen on the basis of the possible<br />

link of the detection of specific odorants to known disease state.<br />

For example, using random presentation, 5 different odors were<br />

presented at 9 different levels of concentration twice to a subject<br />

with End Stage Renal Disease (ESRD). Blanks were presented 9<br />

times also. Total administration time was 38 minutes. The subject<br />

was 100% consistent <strong>for</strong> isoamyl acetate, 89% percent consistent<br />

<strong>for</strong> P-cresol, and 44% consistent <strong>for</strong> vanillin and blanks, i.e. the<br />

consistency was less than chance suggesting that the subject<br />

was guessing. This result is consistent with the prediction that<br />

P-cresol may block the detection of vanillin in ESRD patients.<br />

Acknowledgements: William H Wheeler Foundation<br />

#P12 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Compensation Gone Awry: Conditions Inducing Regional<br />

Oral Sensory Loss May Elevate Obesity Risk<br />

Derek J Snyder 1,2 , Linda M Bartoshuk 1,2<br />

1<br />

Center <strong>for</strong> Smell and Taste, University of Florida Gainesville,<br />

FL, USA, 2 Community Dentistry, University of Florida<br />

Gainesville, FL, USA<br />

Oral afferent nerves innervate portions of the mouth, convey<br />

unique arrays of in<strong>for</strong>mation, and take different paths to<br />

the brain. As such, certain health conditions predict specific<br />

patterns of regional oral sensory loss: Severe childhood ear<br />

infections (otitis media, OM) damage the chorda tympani (CT)<br />

and block anterior taste cues, while tonsillectomy damages the<br />

glossopharyngeal nerve (IX) and blocks posterior taste/tactile<br />

cues. These local effects disinhibit intact oral sensations; <strong>for</strong><br />

example, CT block elevates IX, trigeminal, and whole-mouth<br />

intensity, particularly in supertasters of 6-n-propylthiouracil.<br />

Although this compensatory mechanism often mitigates realworld<br />

(i.e., whole-mouth) perceptual deficits, our recent data<br />

implicate it in a more subtle trend toward long-term obesity<br />

risk. In a laboratory study (N = 301), individuals with medical<br />

histories indicating either CT damage (i.e., OM) or IX damage<br />

(i.e., tonsillectomy) show elevated whole-mouth taste, oral<br />

burn and viscosity, and retronasal olfaction (RO). In survey<br />

data from a larger sample (N = 6584), such individuals also<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

33


show increased high-fat food avidity and body mass. Consistent<br />

with reports linking RO to oral sensation, those with both OM<br />

and tonsillectomy show reduced whole-mouth taste and RO,<br />

revealing compensatory limits following extensive loss. While<br />

this model remains exploratory – it requires verification in a<br />

single sample, and shifts in obesity risk with whole-mouth gain<br />

vs. loss remain to be seen – relative intensity changes among<br />

flavor components appear sufficiently robust to influence<br />

high-fat intake. Overall, oral disinhibition sustains whole-mouth<br />

taste and flavor perception by moderating the impact of limited<br />

spatial loss, but the strength of this effect may shape long-term<br />

food choice and dietary health. Acknowledgements: NIDCD<br />

(DC 00283)<br />

#P14 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Atherosclerosis and Decline in Odor Identification<br />

Carla R. Schubert 1 , Karen J. Cruickshanks 1,2 , Mary E. Fischer 1 ,<br />

Guan-Hua Huang 3 , Barbara E. Klein 1 , Ronald Klein 1 ,<br />

James S. Pankow 4 , Nathan Pankratz 5 , Alex Pinto 1<br />

1<br />

University of Wisconsin/Department of Ophthalmology & Visual<br />

<strong>Sciences</strong> Madison, WI, USA, 2 University of Wisconsin/Department<br />

of Population Health <strong>Sciences</strong> Madison, WI, USA, 3 National Chiao<br />

Tung University/Institute of Statistics Hsinchu, Taiwan, 4 University<br />

of Minnesota/Division of Epidemiology and Community Health<br />

Minneapolis, MN, USA, 5 University of Minnesota/Laboratory<br />

Medicine and Pathology Minneapolis, MN, USA<br />

#P13 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Head Trauma, Taste Damage and Weight Gain<br />

Linda M. Bartoshuk 1 , Susan Marino 2 , Derek J. Snyder 1 ,<br />

Jennifer J. Stamps 1<br />

1<br />

University of Florida Gainesville, FL, USA, 2 University of Minnesota<br />

Minneapolis, MN, USA<br />

Severe brain injuries associated with weight gain have been<br />

reported in children (Jourdan et al. 2012; Norwood et al. 2010)<br />

and adults (Henson et al. 1993); one source of the weight gain<br />

is believed to be metabolic (e.g., hypothalamic dysfunction).<br />

Head trauma has long been known to affect taste (Sumner 1967;<br />

Costanzo and Zasler 1991; Solomon et al. 1991). Our recent<br />

data suggest that taste damage from otitis media or tonsillectomy<br />

is associated with enhanced palatability of energy dense foods<br />

and weight gain (see Snyder & Bartoshuk poster). We presented<br />

a model suggesting that damage to taste nerves VII or IX could<br />

intensify non-taste oral sensations centrally possibly altering<br />

palatability (Bartoshuk et al. 2012). Of special interest, the<br />

consequences of damage to either VII or IX depended on the<br />

status of the other nerve. Central intensifications only occurred<br />

when the damage was restricted to one nerve. The present<br />

study extends these conclusions to taste damage resulting from<br />

head trauma. Subjects were healthy academics who reported<br />

they had experienced a concussion, loss of consciousness<br />

or loss of memory from a head injury. In a questionnaire<br />

study (N=3807), weight was significantly elevated and so was<br />

preference <strong>for</strong> high fat foods controlling <strong>for</strong> age and sex. Spatial<br />

taste testing (N=287) revealed significant loss of taste at VII<br />

with intensifications in flavor and oral touch (e.g., fat) in those<br />

subjects with intact IX. A third population is of special interest:<br />

athletes in contact sports like boxing or football who tend to gain<br />

weight in retirement. We suggest that food preferences should be<br />

examined in head trauma to determine how much weight gain<br />

can be attributed to taste damage with resulting sensory and<br />

palatability alterations. Acknowledgements: DC283, DC8613<br />

and DC8620<br />

Olfactory impairment is common in older adults although<br />

awareness of impairment is low, suggesting that olfactory<br />

function may decline slowly over time. We evaluated factors<br />

associated with a 5-year decline in odor identification in the<br />

Beaver Dam Offspring Study, a longitudinal cohort study of<br />

adults aged 21-84 years at baseline (BOSS1; 2005-2008). The<br />

8-odorant San Diego Odor Identification Test (SDOIT) was<br />

administered at the baseline and 5-year follow-up (BOSS2;<br />

2010-2013) examinations. Decline in odor identification was<br />

defined as a decrease in SDOIT score ≥2 from BOSS1 to<br />

BOSS2; no change was defined as a difference between BOSS1<br />

and BOSS2 scores of ≤1. In preliminary analyses of the first<br />

2195 participants with SDOIT data at BOSS1 and BOSS2,<br />

3.1% had a decline in SDOIT score. Those with a decline in<br />

odor identification were more likely to be older (Odds Ratio<br />

(OR)=1.58, 95% Confidence Interval(CI)=1.40, 1.79, per 5<br />

years of age) than those with no change in score. In age- and<br />

sex-adjusted models, baseline current smoking (OR=2.18,<br />

95%CI=1.08, 4.37, vs never), carotid artery intima media<br />

thickness(IMT)(OR=1.17, 95%CI=1.01, 1.36, per 0.1mm),<br />

number of carotid artery sites (range 0-6) with plaque (OR=1.39,<br />

95% CI=1.13, 1.70) and report of a head injury between BOSS1<br />

and BOSS2 (OR=3.07, 95%CI=1.23, 7.64) were associated with<br />

an increased risk <strong>for</strong> decline in SDOIT score. In a multivariable<br />

model adjusting <strong>for</strong> age, sex and head injury, the number of<br />

carotid artery sites with plaque was an independent predictor of<br />

decline in SDOIT score (OR=1.37, 95%CI=1.11, 1.68). Smoking<br />

and IMT were not significant in models including plaque. These<br />

preliminary findings suggest atherosclerosis may be a contributor<br />

to, or marker of, the decline in olfactory function seen with<br />

aging. Acknowledgements: The project described was supported<br />

by R01AG021917 from the National Institute on Aging,<br />

National Eye Institute, and National Institute on Deafness and<br />

Other Communication Disorders. The content is solely the<br />

responsibility of the authors and does not necessarily reflect the<br />

official views of the National Institute on Aging or the National<br />

Institutes of Health.<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

34


#P15 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Measures of smell function in youth with autism<br />

E. Leslie Cameron 1 , Richard L. Doty 2 , Shereen J. Cohen 3 ,<br />

Karen R. Dobkins 3<br />

1<br />

Department of Psychological Science, Carthage College Kenosha, WI,<br />

USA, 2 Smell and Taste Center, Department of Otorhinolaryngology:<br />

Head and Neck Surgery Philadelphia, PA, USA, 3 Department of<br />

Psychology San Diego, CA, USA<br />

Autism spectrum disorders (ASD) are pervasive developmental<br />

disorders characterized by deficits in social, communicative,<br />

and emotional behaviors. In addition to these hallmarks, there<br />

is evidence <strong>for</strong> atypical sensory processing. In particular, there<br />

is the suggestion of atypicalities in chemosensation, tested<br />

with either questionnaires or direct smell tests. Here we used<br />

both measures concomitantly. Sixteen youth with ASD (mean<br />

15.3 yrs, 4 girls) and 16 typically developing youth (mean 14.3<br />

yrs, 7 girls) participated. Self-reported sensory processing was<br />

measured with the Adolescent/Adult Sensory Profile. This<br />

60-item questionnaire includes items related to the five major<br />

senses and conceptually arranged into four categories – “low<br />

registration” (i.e. low sensitivity to stimuli), “sensory sensitivity”<br />

(i.e. high sensitivity to stimuli), “sensation seeking”, and<br />

“sensation avoiding”. Smell function was measured with the<br />

pediatric Smell Wheel (Cameron & Doty, 2012). This scratch<br />

and sniff test measures the ability to identify 11 common<br />

odors using a 4-alternative <strong>for</strong>ced-choice procedure using<br />

pictures and words to reduce cognitive load. Participants also<br />

rated the pleasantness of each odor. Youth with ASD scored<br />

significantly higher on “low registration”, “sensory sensitivity”<br />

and “sensation avoidance”, and significantly lower on “sensation<br />

seeking” (p


suggest that latency of the N1 OERP component during retrieval<br />

of an odor recognition memory task may be a useful measure <strong>for</strong><br />

examining the negative effects of high waist to hip ratios in those<br />

genetically at risk <strong>for</strong> AD. Acknowledgements: Supported by<br />

NIH grant # DC002064-14 from the NIDCD and A6004085-25<br />

from the NIA. We thank Paul Gilbert <strong>for</strong> his statistical expertise<br />

and Derek Snyder, Jessica Bartholow, Roberto Zamora, Ariana<br />

Stickel, Kyle Sigel, Jean-Loup Bitterlin, Kristina Constant,<br />

and Sanae Okuzawa <strong>for</strong> helping with data collection, entry<br />

and analysis.<br />

#P19 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Odor perception and cerebral odor processing in adults with<br />

autism spectrum condition.<br />

Luzie K. Koehler 1 , Cornelia Hummel 1 , Katja Albertowski 2 ,<br />

Veit Roeßner 2 , Thomas Hummel 1<br />

1<br />

Smell & Taste Clinic, Department of Otorhinolaryngology,<br />

University of Dresden Medical School Dresden, Germany,<br />

2<br />

Department of Children and Youth Psychiatry, University of<br />

Dresden Medical School Dresden, Germany<br />

#P18 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Adverse effect of non-occupational atmospheric exposure to<br />

manganese on peripheral and central olfactory function<br />

Marco Guarneros 1,2 , Nahum Ortiz-Romo 1 , Mireya Alcaraz-Zubeldía 3 ,<br />

Matthias Laska 4 , René Drucker-Colín 1 , Robyn Hudson 5<br />

1<br />

Instituto de Fisiología Celular, Universidad Nacional Autonoma<br />

de Mexico (UNAM) Mexico City, Mexico, 2 Posgrado en Ciencias<br />

Biológicas, UNAM Mexico City, Mexico, 3 Instituto Nacional de<br />

Neurología y Neurocirugía Mexico City, Mexico, 4 Department of<br />

Physics, Chemistry and Biology, Linköping University Linköping,<br />

Sweden, 5 Instituto de Investigaciones Biomédicas, UNAM Mexico<br />

City, Mexico<br />

Manganese (Mn), although an essential trace element, is of<br />

growing concern as a toxic air pollutant. It is not only readily<br />

transported from the olfactory epithelium (OE) to the olfactory<br />

bulb but, unlike other metals, is also transported transsynaptically<br />

to structures deep within the brain. However, to our knowledge<br />

no in<strong>for</strong>mation is available on the possible effect of nonoccupational<br />

exposure to Mn on olfactory function. We there<strong>for</strong>e<br />

investigated peripheral and central olfactory per<strong>for</strong>mance in a<br />

non-occupationally Mn-exposed population. Using the Sniffin’<br />

sticks test battery, we compared the olfactory per<strong>for</strong>mance of<br />

two groups (n = 30/group) from a Mn mining district in central<br />

Mexico: an exposed group living


objective olfactory function in CD patients. 31 CD patients in<br />

active disease, 27 CD patients in inactive disease and a control<br />

sample of 35 age and sex matched healthy volunteers were<br />

included in the study. Subjective testing was applied using the<br />

Sniffin Sticks Test. For olfactory testing olfactory event-related<br />

potentials (OERPs) were obtained by a four channel olfactometer<br />

using phenyl ethyl alcohol (PEA) and hydrogen sulfide (H 2<br />

S).<br />

Chemosomatosensory event-related potentials (CSSERPs) were<br />

obtained using carbon dioxid (CO 2<br />

) as a control stimulus. The<br />

analysis of variance revealed significant higher hedonic ratings<br />

<strong>for</strong> pleasant odorants in CD patients compared to healthy<br />

controls. Furthermore a trend to a lower threshold in CD<br />

patients in inactive disease occurred. In the objective olfactory<br />

testing CD patients showed lower base-to peak-amplitudes<br />

(N1) and lower peak-to-peak-amplitudes (P1-N1). Additionally<br />

the OERPs showed significant shorter N1- and P2-latencies<br />

following the unpleasant odorant (H 2<br />

S) in the right nostril<br />

<strong>for</strong> inactive disease compared to healthy controls. Our results<br />

demonstrate specific abnormities of olfactory perception in<br />

CD patients, which could be interpreted as a natural regulation<br />

against malnutrition. This points out the significance of the<br />

interaction between gastrointestinal functions and olfactory<br />

perception. Acknowledgements: ELAN-Application / FAU,<br />

in part by Neurotrition Project, FAU Emerging Fields Initiative.<br />

#P21 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Effects of Parkinson’s Disease on the Global Field Power<br />

of Olfactory Event-Related Potentials<br />

Allen Osman 1 , Ellen Carson 1 , Ian Pawasarat 1 , Fidias E.<br />

Leon-Sarmiento 1 , Jonathan Silas 2 , Richard L. Doty 1<br />

1<br />

Smell and Taste Center, University of Pennsylvania Philadelphia, PA,<br />

USA, 2 Department of Psychology, Roehampton University London,<br />

United Kingdom<br />

Aim: Measurement of event-related brain potentials in<br />

response to olfactory stimulation (OERPs) most often involves<br />

the amplitude and/or latency of specific components. These<br />

components however can be difficult to discern in OERPs<br />

from a single individual, especially one who is hyposmic. To<br />

circumvent this problem, we explored the use of an alternative<br />

measure that reflects the overall strength of an OERP: Its Global<br />

Field Power (GFP). Methods: The subjects were 20 patients<br />

with early-stage Parkinson’s disease (PD) and 20 matched<br />

healthy controls. Stimuli consisted of 200 msec. presentations of<br />

hydrogen sulfide delivered via a continuous-flow olfactometer.<br />

OERPs were recorded at 40 electrode sites across the scalp and<br />

combined to <strong>for</strong>m a single GFP trace based on their standard<br />

deviation at each time-point. Results: It was possible to obtain<br />

GFP measures <strong>for</strong> almost all subjects. The magnitude of change<br />

in GFP following odorant presentation, i.e. the overall strength<br />

of the OERP response, was smaller <strong>for</strong> PD than control subjects.<br />

Conclusions: While previous work has found effects of PD on<br />

OERP latency, the present study shows that there is an effect<br />

also on amplitude. More generally, this study demonstrates<br />

that a fully objective OERP measure is sensitive to olfactory<br />

function and can be obtained readily from single individuals. As<br />

such, OERP GFP may be of use in the clinic and well suited <strong>for</strong><br />

standard statistical analyses. Acknowledgements: Supported by<br />

USAMRAA W81XWH-09-1-0467.<br />

#P22 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Sensory Modality Influences Episodic Metamemory<br />

Accuracy in Healthy Aging and Alzheimer’s Disease<br />

Jacquelyn Szajer 1 , Claire Murphy 1,2<br />

1<br />

San Diego State University San Diego, CA, USA,<br />

2<br />

University of Cali<strong>for</strong>nia San Diego San Diego, CA, USA<br />

Episodic memory is commonly known to decline in both healthy<br />

aging and Alzheimer’s disease (AD), however, these declines<br />

are heterogeneous. One factor shown to influence declines in<br />

episodic memory is metamemory, which is known to play an<br />

essential role in the strategies used <strong>for</strong> the encoding and retrieval<br />

of in<strong>for</strong>mation, as well as the control of memory output (Pannu<br />

& Kaszniak, 2005). Prior research has shown that memory-based<br />

olfactory tests are useful in gauging cognitive decline in aging<br />

and AD (Murphy, Nordin, & Jinich, 1999), and the current<br />

study aimed to examine the utility of these tests in assessing<br />

declines in metamemory accuracy. Using a sample of 109<br />

older adults with mild to moderate Alzheimer’s disease and 97<br />

healthy controls, the effect of diagnosis and sensory modality<br />

on metamemory accuracy (operationally defined as confidence<br />

in the accuracy of responses at retrieval) was analyzed using an<br />

episodic recognition memory task comprised of both olfactory<br />

and visual stimuli. Results indicated that both diagnosis and<br />

modality had main effects on confidence accuracy (p


#P23 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

#P24 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Primary Olfactory Cortex is affected in Alzheimer’s<br />

Disease and Mild Cognitively Impaired Patients:<br />

A neuroimaging study<br />

Megha Vasavada 1 , Jianli Wang 1 , Xiaoyu Sun 1 , Christopher<br />

Weitekamp 1 , Paul Eslinger 2 , Prasanna Karunanayaka 1 ,<br />

Sarah Ryan 1 , Qing Yang 1,3<br />

1<br />

Radiology, Penn State College of Medicine Hershey, PA, USA,<br />

2<br />

Neurology, Penn State College of Medicine Hershey, PA, USA,<br />

3<br />

Neurosurgery, Penn State College of Medicine Hershey, PA, USA<br />

Alzheimer’s Disease is a neurodegenerative disorder affecting<br />

5.4 million Americans and it is the 6 th leading cause of death.<br />

Diagnosis of the disease is made when the pathology has<br />

progressed to the neocortex and the effectiveness of drug<br />

intervention is unlikely. There<strong>for</strong>e, early diagnosis is key in<br />

understanding the disease progression and unlocking a cure. It<br />

has been shown that the pathology of AD (amyloid beta plaques<br />

(Ab) and neurofibrillary tangles (NFT)) are first found in areas<br />

involved in olfaction. Decreased sense of smell is seen in the<br />

earliest stages of AD and in Mild Cognitive Impaired (MCI)<br />

patients. In this study we used olfactory functional Magnetic<br />

Resonance Imaging (fMRI) and volumetric MRI to examine the<br />

relationship between the functional deficit and the pathological<br />

changes (atrophy) in the primary olfactory cortex (POC) and<br />

in the hippocampus in 23 cognitively normal controls (NC), 19<br />

MCI, and 15 AD subjects. The volumetric data shows that the<br />

volume of the POC is significantly different between the three<br />

groups (p


#P25 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

#P26 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Differences in Odor Identification between Alzheimer’s and<br />

Parkinson’s Patients<br />

Jennifer J Stamps 1 , Kenneth M Heilman 2 , Linda M Bartoshuk 1<br />

1<br />

University of Florida Center <strong>for</strong> Smell and Taste Gainesville, FL, USA,<br />

2<br />

University of Florida Department of Neurology Gainesville, FL, USA<br />

Whereas odor detection threshold is more impaired in<br />

patients with Parkinson’s disease (PD) than Alzheimer’s disease<br />

(AD), odor identification appears more impaired in AD than<br />

PD (Rahayel et al.’12). Failure to identify sensory stimuli can<br />

be induced by a sensory deficit (anosmia, failure to detect), a<br />

perceptual recognition deficit (apperceptive agnosia, ability<br />

to detect the odor without the ability to correctly describe the<br />

odor, and denies correct identification), or a failure of percept<br />

to access lexical semantic networks (anomia, inability to name<br />

with preserved ability to correctly describe the odor, select<br />

the name from choices or agree when told the correct name).<br />

The purpose of this study was to better understand the type of<br />

olfactory identification error exhibited in 22 participants with<br />

AD, 23 with PD and 45 matched controls (HC) during an openchoice<br />

olfactory task (NO) with 23 food items. The percentages<br />

of correct odor identification of AD (39%) and PD (42%)<br />

participants were not different, but both were significantly lower<br />

than HC (83%, p


#P28 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Taste and Odor Convergence in the Nucleus of the<br />

Solitary Tract of Awake, Behaving Rats.<br />

Olga D Escanilla, Patricia M Di Lorenzo<br />

Binghamton University/Psychology Department Binghamton,<br />

NY, USA<br />

Although many studies have shown the importance of gustatory<br />

and olfactory interactions in flavor perception, very little is<br />

known about multisensory interaction in the initial stages of<br />

gustatory processing. Previously, it has been shown that a subset<br />

of cells in the nucleus of the solitary tract (NTS; Van Buskirk &<br />

Erickson, Brain Res.,135(2):287-303, 1977) and the parabrachial<br />

nucleus of the pons (PbN; Di Lorenzo & Garcia, Brain Res<br />

Bull.,15(6):673-6, 1985) in rats respond to both taste and<br />

olfactory stimuli. Here, we studied whether taste cells in the NTS<br />

of the awake, behaving animal could also respond to olfactory<br />

stimuli. Rats were surgically implanted with a microwire bundle<br />

into the NTS and allowed to recover. Rats were mildly water<br />

deprived and placed in an experimental chamber containing<br />

a lick spout <strong>for</strong> taste stimulus delivery and an odor port <strong>for</strong><br />

olfactory stimulus delivery. Tastants (0.1 M NaCl, 0.1 M sucrose,<br />

0.01 M citric acid, 0.0001 M quinine and artificial saliva) were<br />

delivered <strong>for</strong> 5 consecutive licks interspersed with 5 licks of<br />

artificial saliva rinse delivered on a variable ratio 5 schedule. All<br />

taste stimuli were presented both with and without an odorant<br />

in separate trials. Odorants were 1 Pa n-amyl acetate, 1 Pa<br />

acetic acid and air. Of the 31 cells recorded thus far, 74% were<br />

taste responsive, and 23% were odor responsive. There were<br />

no odor-responsive cells that were not also taste-responsive.<br />

When odorants are paired with taste stimuli, 96% of the 31 cells<br />

recorded showed either suppression or enhancement of taste<br />

responses. In addition, a small group of non-taste- or odorantresponsive<br />

cells (n = 8) responded when presented with a paired<br />

odorant-tastant stimulus. These results suggest that multisensory<br />

processing occurs at the initial stages of gustatory processing.<br />

Acknowledgements: Supported by NIDCD grant RO1DC006914<br />

to PMD.<br />

#P29 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Enhancement of odor intensity and hedonics by taste:<br />

roles of nutritive taste and congruency<br />

Tomomi Fujimaru, Juyun Lim<br />

Oregon State University Corvallis, OR, USA<br />

We have previously shown that sucrose enhances the perceived<br />

intensities of congruent retronasal odors, whereas caffeine and<br />

citric acid do not. The present study is designed to investigate<br />

whether saltiness and umami can also enhance retronasal<br />

odors. In addition, given our previous finding that tasteodor<br />

congruency plays an important role in retronasal odor<br />

referral to the mouth, we tested the roles of congruency in the<br />

enhancement of odor intensities and hedonics. Tomato and<br />

chicken odors were presented alone or with NaCl, KCl, MSG,<br />

MPG, or caffeine. Using a sip-and-spit procedure, Ss rated 1) the<br />

degree of liking/disliking of flavor on the LHS; 2) the intensities<br />

of saltiness, savoriness, bitterness and specific odor on the gLMS;<br />

and 3) the degree of taste-odor congruency on a VAS. The result<br />

showed that both salty (NaCl) and umami (MSG, MPG) tastes<br />

significantly enhanced the perceived intensities of tomato and<br />

chicken odors (Tukey test, p


TAS2R38 genotype did not associate with the intensity of<br />

these sensations. When individuals were split by genotype, the<br />

strength of the PROP-capsaicin and PROP-sucrose relationships<br />

increased substantially within the groups of homozygous<br />

individuals. Collectively, this suggests PROP bitterness is a<br />

confounded phenotype that captures both genetic variation<br />

specific to N-C=S compounds and overall orosensory response.<br />

Acknowledgements: Supported by funds from the Pennsylvania<br />

State University and NIH grant DC0010904.<br />

#P32 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Understanding Valence: the Neurobiology of Appetitive<br />

and Aversive Odor-Taste Learning in Rats<br />

SiWei Luo, Matthew Einhorn, Kim M. Gruver, Nana Fujiwara,<br />

Thomas A. Cleland<br />

Cornell University/Psychology Ithaca, NY, USA<br />

#P31 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

The Effect of Retronasal Odor on Ratings of Sweetness<br />

and Bitterness<br />

Tomoyuki Isogai 1,2 , Paul Wise 2<br />

1<br />

Milk science Institute, Megmilk Snow Brand Co.,Ltd. Kawagoe, Japan,<br />

2<br />

Monell Chemical Senses Center Philadelphia, PA, USA<br />

We report the first in a planned series of experiments on how<br />

odors affect rated bitterness, a potential flavor interaction<br />

that has received less attention than those with other taste<br />

sub-modalities (viz., sweet, sour, salty, and savory). These<br />

experiments used a computer-controlled olfactometer-gustometer<br />

to simultaneously present taste solutions and odorized air to<br />

the mouth. One experiment replicated a known effect, viz.<br />

enhancement of sweetness by a fruity-smelling ester, using the<br />

new olfactometer-gustometer. A second experiment examined<br />

the effect of a “burnt” odor on bitter intensity. Subjects included<br />

12 healthy men and women, aged 18 to 65. Taste solutions<br />

included five concentrations each of sucrose (sweet) or sucrose<br />

octaacetate (SOA, bitter). Odorants included four concentrations<br />

each of ethyl hexanoate (sweet, pineapple) or isovaleraldehyde<br />

(burnt meat). Subjects held solutions in the mouth <strong>for</strong> several<br />

seconds, and rated the strength of both taste and odor sensations<br />

be<strong>for</strong>e expectorating. Subjects rated intensity using the general<br />

labeled magnitude scale (gLMS). Ratings of sweetness and<br />

bitterness increased with the concentrations of sucrose and<br />

SOA, respectively. This expected dose-response relationship<br />

supported the validity of the ratings. Further, consistent with<br />

published results, ethyl hexanoate significantly enhanced the<br />

rated sweetness of sucrose solutions. Thus, the current method<br />

of automated testing proved sensitive to known enhancement<br />

effects. Finally, isovaleraldehyde significantly enhanced the<br />

bitterness of SOA solutions, demonstrating that odors can<br />

enhance bitter taste. In conclusion, the method per<strong>for</strong>med well,<br />

and modulation of bitter taste by retronasal odors seems like a<br />

promising area <strong>for</strong> further investigation.<br />

Appetitive and aversive conditioning both exert effects on<br />

perceptual learning, but are qualitatively distinct; <strong>for</strong> example,<br />

they evoke activity in different brain regions. Here, we present<br />

an olfactory conditioning paradigm capable of evoking both<br />

<strong>for</strong>ms of conditioning with minimal procedural differences,<br />

enabling conditioning effects to be directly compared using<br />

identical behavioral and physiological analyses. Male Long-<br />

Evans rats (Rattus norvegicus) served as subjects <strong>for</strong> this study.<br />

Using implanted intra-oral cannulae, appetitive or aversive<br />

tastants (0.20% saccharin or 0.02 M quinine) were directly<br />

infused into rats’ mouths and paired (or backward-paired) with<br />

an odor conditioned stimulus over 3 days such that training<br />

procedures differed only in the valence of the tastant. Rats<br />

displayed appropriate anticipatory behaviors (rapid mouth<br />

movements, tongue protrusions, gaping) in response to odors<br />

predicting tastant infusions, indicating that rats learned the<br />

association between odors and tastants. Aversively-conditioned<br />

rats also appeared to exhibit broader generalization to similar<br />

odorants than did the appetitive and control groups. Immediateearly<br />

gene (Egr-1 and c-Fos) expression was measured in<br />

olfaction-, valence-, and conditioning-associated brain regions.<br />

IEG expression was higher in the main olfactory bulb, anterior<br />

olfactory nucleus, piri<strong>for</strong>m cortex, and orbitofrontal cortex in<br />

aversively-conditioned rats compared with appetitive and control<br />

groups. These results provide a foundation <strong>for</strong> studies of learning<br />

and plasticity in which appetitive and aversive associations can<br />

be mechanistically compared in a common neural network.<br />

Acknowledgements: Liu Memorial Award, Sigma Xi Research<br />

Grant, Cornell University SAGE Fellowship<br />

#P33 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Pou2f3/Skn-1a is involved in the differentiation of multiple<br />

types of Trpm5-expressing chemosensory cells<br />

Makoto Ohmoto 1 , Tatsuya Yamaguchi 2 , Keiko Abe 3 ,<br />

Alexander A. Bachmanov 1 , Junji Hirota 2 , Ichiro Matsumoto 1<br />

1<br />

Monell Chemical Senses Center Philadelphia, PA, USA, 2 Tokyo<br />

Institute of Technology, Department of Bioengineering Yokohama,<br />

Japan, 3 The University of Tokyo, Department of Applied Biological<br />

Chemistry Tokyo, Japan<br />

POSTER PRESENTATIONS<br />

A homeodomain transcription factor Pou2f3 (also known as<br />

Skn-1a) is specifically expressed in sweet, umami, and bitter taste<br />

receptor cells in taste buds, and is necessary <strong>for</strong> their functional<br />

differentiation 1 . Recent studies have shown that taste receptors<br />

and/or signaling molecules indispensable <strong>for</strong> sweet, umami,<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

41


and/or bitter taste are expressed not only in gustatory tissues<br />

but also in the intestinal and respiratory epithelia. In this study,<br />

we examined the expression of Pou2f3 in the extra-oral epithelial<br />

chemosensory cells and the impact of Pou2f3 knockout on the<br />

cells. In the intestinal epithelium, the expression of Pou2f3<br />

was observed in the tuft/brush cells that express Plcb2 and<br />

Trpm5 and participate in opioid secretion by chemosensing.<br />

Pou2f3-deficient mice lacked the expression of Plcb2, Trpm5,<br />

and other tuft/brush cell marker genes, but they still expressed<br />

characteristric genes of enteroendocrine cells. In the respiratory<br />

epithelium of the nasal cavity, the expression of Pou2f3 was<br />

observed in the solitary chemosensory cells (SCCs) that express<br />

Tas1r3, Tas2rs, Gnat3, Plcb2, and Trpm5. The expression of<br />

all these genes was lost in the Pou2f3-defficient mice. We also<br />

found Pou2f3 expression in a subset of microvillus cells in the<br />

main olfactory epithelium (MOE) where Trpm5 but not Plcb2,<br />

Ggust, or taste receptors were expressed. Pou2f3-deficient mice<br />

exhibited the lack of Trpm5 expression in the MOE. Taken<br />

together, these data demonstrate that Pou2f3 expression is<br />

associated with the expression of Trpm5 in multiple types<br />

of chemosensory cells, and suggest that Pou2f3 is a master<br />

regulator of differentiation of Trpm5-expressing chemosensory<br />

cells in digestive and respiratory epithelia. 1 Nat. Neurosci.<br />

14, 685-687 (2011)<br />

#P34 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Aronia Berry Juice Sensory Analysis by Harvest Time and<br />

Oral Sensory Phenotype<br />

Jeeha Park 1 , Shristi Rawal 1 , Mark H Brand 2 , Shelley Durocher 2 ,<br />

Mastaneh Sharafi 1 , Valerie B Duffy 1<br />

1<br />

University of Connecticut/Allied Health <strong>Sciences</strong> Storrs, CT, USA,<br />

2<br />

University of Connecticut/Plant Science and Landscape Architecture<br />

Storrs, CT, USA<br />

Aronia berries (chokeberries) have very high levels of healthpromoting<br />

antioxidants yet can cause a “choking” sensation<br />

due to bitterness and astringency. We aimed to describe oral<br />

sensations and palatability of aronia juice, including variation<br />

by harvest time and oral sensory phenotype. Ripe aronia berries<br />

were harvested at 7 time points and juiced <strong>for</strong> oral sampling<br />

by 50 adults who were phenotyped <strong>for</strong> chemosensory function<br />

by the NHANES protocol, olfactometer, and propylthiouracil<br />

(PROP) bitterness. The adults reported quality intensities of<br />

prototypical tastes, oral chemesthetic compounds, foods, berry<br />

juice and non-oral stimuli, which served as sensory standards.<br />

The juice qualities correlated with alum astringency, quinine<br />

bitterness, and citric acid sourness but, unlike apple or grapefruit<br />

juices, did not correlate with sucrose sweetness. The average<br />

berry juice response was weakly dislike (range—v. strongly<br />

dislike to above moderately like). Astringency was the strongest<br />

sensation, yet sweetness was the primary driver of liking in<br />

multiple regression analysis. Those who liked the juice reported<br />

a greater balance between astringency and either sourness or<br />

sweetness. The juices were more sweet and less sour/astringent<br />

in the middle versus first harvest time, corresponding to greater<br />

acceptance. Classifying by PROP relative to quinine bitterness<br />

identified adults who differed in intensities of prototypical oral<br />

stimuli but not in intensities of odors. Those who perceived more<br />

PROP relative to quinine bitterness also reported the juice as less<br />

sour, less balanced between astringency and sourness, and more<br />

disliked than those who perceived high bitterness from PROP<br />

and quinine. Further study aims to identify ways to maximize<br />

the palatability of aronia juice, individualized to taste phenotype.<br />

Acknowledgements: USDA/Hatch and USDA NE SARE<br />

#P35 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Temperature of served water can influence sensory perception<br />

and acceptance of subsequent food<br />

Han-Seok Seo 1 , Pauline Mony 1,2 , Tonya Tokar 1 , Peggy Pang 1 ,<br />

Alexandra Fiegel 1 , Jean-François Meullenet 1<br />

1<br />

University of Arkansas/Food Science Fayetteville, AR, USA,<br />

2<br />

French National School of Agricultural Science and Engineering in<br />

Toulouse Castanet Tolosan Cedex, France<br />

The cross-cultural difference in meal pattern exists in the typical<br />

temperature of water served with meals. For example, North<br />

American people, as a whole, are used to drinking iced water/<br />

beverages, while Asian or European people show a preference<br />

<strong>for</strong> hot water/tea or room temperature water, respectively. It has<br />

been proven that food perception and acceptance are affected<br />

by oral temperature, as well as by serving temperature of food.<br />

Based on the fact that the iced or hot water served with meals<br />

can modulate the oral temperature, the present study aimed to<br />

determine if the temperature of served water can modulate the<br />

sensory perception of foods subsequently consumed. Following<br />

a mouth rinse with water served at 4, 20, and 50 °C <strong>for</strong> 5 s,<br />

two types of food: dark chocolate or cheddar cheese were<br />

evaluated in terms of sensory intensity and overall liking. For<br />

the dark chocolate, the intensity ratings <strong>for</strong> sweetness, chocolate<br />

flavor, and creaminess were significantly lower when following<br />

water at 4 °C than when following water at either 20 or 50 °C.<br />

However, the effect of water temperature on sensory perception<br />

was not observed with cheddar cheese. In addition, the overall<br />

liking <strong>for</strong> the dark chocolate was significantly lower when<br />

following water at 4 °C than when following water at either 20<br />

or 50 °C. In conclusion, the current study demonstrates new<br />

empirical evidence that the consumption of iced water can<br />

decrease perceived intensities of sweetness, chocolate flavor,<br />

and creaminess <strong>for</strong> subsequently consumed chocolate. Our<br />

findings suggest a possibility that the North American frequent<br />

consumption of iced water/soda may reduce their sensitivity to<br />

sweet tasting stimuli, thereby leading to the preference <strong>for</strong> more<br />

highly sweetened foods. Acknowledgements: This research was<br />

supported by start-up funding from the University of Arkansas<br />

Division of Agriculture to HS SEO.<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

42


#P36 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Stimulus temperature and concentration differentially<br />

influence the gustatory neural code <strong>for</strong> sucrose in the<br />

mouse brain stem<br />

David M. Wilson, Christian H. Lemon<br />

Department of Pharmacological and Physiological Science St. Louis<br />

University School of Medicine St. Louis, MO, USA<br />

of the flickering visual object. These results indicate that<br />

olfaction modulates visual temporal processing at the object<br />

representation level, and provide new insights into the neural<br />

timing of multisensory events. Acknowledgements: National<br />

Basic Research Program of China (2011CB711000), National<br />

Natural Science Foundation of China (31070906), National<br />

Natural Science Foundation of China (31100735), Knowledge<br />

Innovation Program of the Chinese Academy of <strong>Sciences</strong><br />

(KSCX2-EW-BR-4 & KSCX2-YW-R-250)<br />

Taste sensitive neurons throughout the gustatory neuraxis<br />

respond to oral somatosensory stimuli, like temperature. What<br />

is more, human psychophysical studies show that unimodal<br />

oral thermal stimuli can elicit taste perceptions, and increasing<br />

the temperature of a sucrose solution increases its perceived<br />

“sweetness”. These data suggest that temperature may be<br />

modulating the neural signal <strong>for</strong> taste intensity. However, data<br />

on this topic are scarce. To investigate the potential influence<br />

of temperature on neural activity <strong>for</strong> taste intensity we made<br />

extracellular recordings from single neurons in the rostral<br />

nucleus of the solitary tract of anesthetized C57BL/6J mice<br />

during oral application of temperature-and concentration-varied<br />

tastants. Stimuli included purified water and a concentration<br />

series of sucrose (in M): 0.05, 0.1, 0.17, 0.31, and 0.56 presented<br />

“whole mouth” at 17, 22, 30, and 37º C. Preliminary analyses<br />

of 35 neurons show that warming sucrose to 37º C significantly<br />

reduced the response onset latency compared to room<br />

temperature sucrose <strong>for</strong> all concentrations tested (ps Males, p=0.016) and pleasantness<br />

ratings of Na and Na/K salt samples (Males rated > Females,<br />

p=0.017 and 0.001). Gender*Temperature interaction was found<br />

<strong>for</strong> pleasantness of Na/K salt samples (p=0.025). Contrary to<br />

expectations, temperature did not have a significant effect on<br />

perceived taste intensity. Nevertheless, it is suggested that sweet<br />

and salt tastes have greater hedonic value at temperatures other<br />

than ambient and that factors such as changes in texture and oral<br />

exposure time should also be considered as the basis <strong>for</strong> the ‘Ice<br />

cream effect’.<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

43


#P39 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Evidence <strong>for</strong> a Cell Fate Refinement Mechanism in<br />

Olfactory Sensory Neurons<br />

Ishmail Abdus-Saboor, Benjamin Shykind<br />

Weill Cornell Medical College Qatar/ Department of Cell and<br />

Developmental Biology Doha, Qatar<br />

Olfactory receptors (ORs) number more than 1,000 and comprise<br />

the largest gene family in the mammalian genome. ORs reside<br />

in heterochromatin and selection of one OR and from one<br />

allele is thought to occur stochastically. ORs are expressed both<br />

monogenically and monoallelically in olfactory sensory neurons<br />

(OSNs) and the mechanism that controls their regulation is<br />

largely unknown. Here we describe results <strong>for</strong> mice with a<br />

‘monoclonal’ nose that express one OR M71in 95% of all mature<br />

OSNs. M71 mice suppress expression of endogenous ORs, and<br />

expression of the suppressed ORs is shifted to the immature<br />

layer of the olfactory epithelium. We show that the suppressed<br />

ORs were first selected and then turned off by M71. When we<br />

introduced a second transgene into M71 mice that expressed<br />

another OR in most mature OSNs, OSNs uncharacteristically<br />

expressed both of the ORs. We hypothesize that unresolved OR<br />

competition compromised the neuron’s ability to express only<br />

one receptor. We further show that suppression of endogenous<br />

ORs by M71 is not reversible, and that M71 does not need<br />

to be continuously expressed <strong>for</strong> endogenous ORs to remain<br />

suppressed. In these experiments, we have engineered OSNs<br />

to express more than one OR in an OSN at a time, which is<br />

normally a low probability event. We have shown that when<br />

this event arises, a secondary refinement pathway is invoked<br />

that turns off one OR to maintain singular expression. We<br />

thus provide compelling evidence <strong>for</strong> a new paradigm of OR<br />

regulation: post-selection shut down, which we hypothesize occurs<br />

through a yet-to-be uncovered competitive mechanism.<br />

#P40 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Odorant Receptor Dependent Spontaneous Firing Rates<br />

Do Not Predict Sensory-evoked Firing Rates in Mouse<br />

Olfactory Sensory Neurons<br />

Timothy Connelly, Agnes Savigner, Minghong Ma<br />

University of Pennsylvania/Department of Neuroscience Philadelphia,<br />

PA, USA<br />

Sensory systems need to tease out stimulation-evoked activity<br />

against a background of spontaneous activity. In the olfactory<br />

system, the odor response profile of an olfactory sensory neuron<br />

(OSN) is dependent on the type of odorant receptor it expresses.<br />

OSNs also exhibit spontaneous activity, which plays a role in<br />

establishing proper synaptic connections and may also increase<br />

the sensitivity of the cells. However, where the spontaneous<br />

activity originates and whether it in<strong>for</strong>ms sensory-evoked activity<br />

remain unclear. We addressed these questions by examining<br />

patch-clamp recordings of genetically labeled mouse OSNs with<br />

the defined odorant receptor M71 (n = 22), I7 (n = 21), SR1<br />

(n = 11), mOR-EG (n = 24) or MOR23 (n = 16) in intact<br />

olfactory epithelia. We show that OSNs expressing different<br />

odorant receptors had significantly different rates of basal<br />

activity. Additionally, OSNs expressing an inactive mutant I7<br />

receptor completely lacked spontaneous activity (n = 34), despite<br />

being able to fire action potentials in response to current injection<br />

(n = 6). This finding strongly suggests that the spontaneous<br />

firing of an OSN originates from the spontaneous activation<br />

of its G-protein coupled odorant receptor. Lastly, we show<br />

that the spontaneous firing rates of selected OSN types do not<br />

correlate with the firing rates evoked by a near-saturating odorant<br />

stimulus. This study reveals that neither the basal activity nor the<br />

receptor type dictates the maximum odorant-evoked activity in<br />

OSNs, which suggests that OSNs expressing the same receptor<br />

type may send distinct in<strong>for</strong>mation to the brain upon odorant<br />

stimulation. Acknowledgements: This work was supported by<br />

R01 grants from NIDCD/NIH (DC006213 and DC011554).<br />

#P41 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

How plastic is the peripheral olfactory system of<br />

Drosophila melanogaster larvae?<br />

MA Grillet, R Petersen, C McCrohan, M Cobb<br />

University of Manchester Manchester, United Kingdom<br />

During their larval stage, fruit flies are exposed to an odourrich<br />

environment, in which they must choose between toxic<br />

and edible substrates. For this they need an efficient olfactory<br />

system with the capacity <strong>for</strong> both short and long term plasticity.<br />

Drosophila larvae possess only 21 paired olfactory sensory<br />

neurons (OSN), most of which express a single olfactory<br />

receptor (OR) type together with the co-receptor Orco.<br />

In<strong>for</strong>mation arising from each OSN is transmitted to a unique<br />

glomerulus in the antennal lobe and then to the mushroom body<br />

via projection neurons. Combinatorial coding in the periphery<br />

allows larvae to detect and discriminate a large number of<br />

odours. Previous studies have characterised the odour-response<br />

profiles of 19 of the 21 OSNs and found that a given OSN’s<br />

response to a given odour is often highly variable (Hoare et<br />

al 2008, 2011). This raised the possibility that the peripheral<br />

olfactory code exhibits plasticity and that this plasticity may<br />

be directly involved in mediating behavioural adaptation and<br />

learning. We are exploiting the UAS-Gal4 system to create single<br />

OR lines in which only one identified OSN is functioning. We<br />

are studying the effect of short and long-term adaptation on the<br />

response of individual OSN classes to a panel of odours using<br />

extracellular electrophysiology and behavioural assays. Our<br />

preliminary results show that short-term adaptation to specific<br />

odours may induce a decrease or an increase in OSN responses<br />

depending on the odour used. We are currently exploring the<br />

effect of longer-term adaptation on the peripheral olfactory code.<br />

Acknowledgements: BBSRC UK grant BB/H009914/1<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

44


#P42 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Characterization of the Iontransporter NKCC1 in the<br />

Field of Chemosensation<br />

Claudia Haering, Janine Wäring, Hanns Hatt<br />

Ruhr-University Bochum, Germany<br />

The olfactory sense is mainly regulated through a cAMPdependent<br />

signaling cascade which leads to a cation influx<br />

and a chloride efflux through the neuronal membrane. This so<br />

called chloride boost during depolarization of olfactory sensory<br />

neurons still remains unclear. In addition, how is the intracellular<br />

chloride concentration achieved in olfactory neurons? Several<br />

publications demonstrated that the chloride concentration<br />

is much higher in olfactory neurons, especially in the knob,<br />

compared to surrounding cells and the mucus. NKCC1 is a<br />

candidate which fits the role of an ion transporter that causes the<br />

high chloride concentration inside olfactory neurons. NKCC1<br />

is a 12 membrane spanning symporter of one sodium, one<br />

potassium and two chloride ions. The expression of NKCC1<br />

is confirmed in the olfactory epithelium of mice, especially<br />

in olfactory neurons. However, the function of NKCC1 is<br />

controversially discussed in literature. In our project we want<br />

to characterize NKCC1 knockout mice, thereby addressing the<br />

question whether NKCC1 is involved in olfaction and olfactory<br />

neurogenesis. On the one hand we are going to use chloride<br />

imaging of acute slices of the olfactory epithelium of knockout<br />

and wild type mice. On the other hand morphological studies<br />

and RNA fluorescence in situ hybridization (RNA FISH)<br />

experiments will give us in<strong>for</strong>mation about the role of NKCC1<br />

in neurogenesis. Our first studies showed differences in the<br />

morphology of the turbinates and the neuronal layer of the<br />

olfactory epithelium of NKCC1 knockout compared to wild type<br />

mice leading to the question whether NKCC1 plays a role in the<br />

continuous replacement of olfactory sensory neurons.<br />

#P43 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Intrinsic electrophysiological property of Kenyon cells<br />

in silkmoths<br />

Shigeki Inoue 1 , Masashi Tabuchi 2 , Kei Nakatani 1 , Ryohei Kanzaki 2<br />

1<br />

University of Tsukuba Tsukuba, Ibaraki, Japan, 2 University of Tokyo<br />

Meguro, Tokyo, Japan<br />

Kenyon cells (KCs) are the component neurons of mushroom<br />

bodies (MBs) of insect brain regions contributing to olfaction<br />

and taste. In these contributions, KCs may have some important<br />

roles in olfactory in<strong>for</strong>mation processing. However, the intrinsic<br />

electrophysiological properties of silkmoth KCs remain<br />

unknown. Here, we use whole-cell patch clamp recording to<br />

elucidate the functional parameters such as voltage-activated<br />

ionic currents of KCs in silkmoth MBs. KCs generated action<br />

potentials in response to depolarizing current injections which<br />

were stepped in 20 pA between 0 and 100 pA, and application of<br />

GABA-receptor blocker, picrotoxin (PTX) abolished inhibitory<br />

synaptic inputs and depolarized resting potential. By using<br />

voltage-clamp technique, we recorded membrane currents<br />

including inward and outward voltage-activated currents.<br />

Pharmacological isolation of KC voltage-activated ionic currents<br />

revealed that KCs express a range of voltage-activated currents,<br />

including transient (I ; activated by voltage step pulses<br />

transient<br />

above -50 mV) and non-activating potassium (I ; activated<br />

sustained<br />

by voltage step pulses above -30 mV), sodium (I ; activated<br />

Na<br />

by voltage step pulses above -50 to -40 mV) and calcium (I ; Ca<br />

activated by voltage step pulse above -60 to -50 mV) currents,<br />

and these potassium currents included calcium-activated<br />

components. Our results consisted with previous research of<br />

cockroach (Demmer and Kloppenburg. 2009) and provided the<br />

first electrophysiological characterization of KCs in silkmoth<br />

MBs and suggested that the intrinsic properties of KCs had<br />

common feature regardless of the insect species. Our experiments<br />

represented an important step toward understanding neural<br />

computation that underlies olfactory in<strong>for</strong>mation processing<br />

in silkmoth.<br />

#P44 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Scaffolding proteins in olfaction<br />

Fabian Jansen 1 , Sabrina Baumgart 1 , Willem Bintig 2 , Benjamin Kalbe 1 ,<br />

Christian Herrmann 3 , David Köster 4 , Sebastian Rasche 1 , Nils Metzler-<br />

Nolte 4 , Marc Spehr 5 , Hanns Hatt 1 , Eva Neuhaus 2<br />

1<br />

Department of Cell Physiology, Faculty <strong>for</strong> Biology and Biotechnology,<br />

Ruhr-Universität Bochum Bochum, Germany, 2 Neuroscience Research<br />

Center, Cluster of Excellence NeuroCure, Charité Universitätsmedizin<br />

Berlin Berlin, Germany, 3 Department of Physical Chemistry I,<br />

Ruhr-Universität Bochum Bochum, Germany, 4 Department of<br />

Inorganic Chemistry I, Ruhr-Universität Bochum Bochum, Germany,<br />

5<br />

Department of Chemosensation, Institute <strong>for</strong> Biology II, RWTH-<br />

Aachen University Aachen, Germany<br />

Mammalian olfactory signaling is dependent on various proteins<br />

and fine-tuned through different processes. Most of the essential<br />

components are already known but the detailed organization<br />

of this complex system has yet to be understood. In our recent<br />

study we identified the multiple PDZ domain protein MUPP1<br />

as a potential scaffolding protein candidate <strong>for</strong> olfactory<br />

signaling. Through large-scale peptide microarrays we could<br />

show numerous interactions between the 13 PDZ domains of<br />

MUPP1 and different murine olfactory receptor C-termini of<br />

various subfamilys. Co-immunoprecipitations validated the<br />

interactions between MUPP1 and the most important signaling<br />

proteins in vivo and led to the assumption that an olfactory<br />

PDZome is organized by the scaffold MUPP1. Furthermore<br />

co-immunoprecipitations in juvenile mice showed no differences<br />

in binding properties in comparison to the adult mice. To<br />

functionally show that olfactory signaling is PDZ-dependent<br />

we established patch clamping of olfactory sensory neurons<br />

in acute slices of transgenic mOR-EG mice. We also created a<br />

small inhibitory peptide which disrupts the interaction between<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

45


the mOR-EG receptor and MUPP1 and injected it through the<br />

patch pipette into the neuron. After uncoupling the interaction<br />

between the mOR-EG receptor and MUPP1 the odor-evoked<br />

current amplitudes were strongly reduced and the adaption was<br />

impaired, whereas a control peptide did not affect olfactory<br />

signaling. In conclusion, we confirmed that an olfactory<br />

signalosome is mediated by MUPP1 in olfactory sensory<br />

neurons and showed that accurate olfactory signaling is a PDZ<br />

dependent mechanism.<br />

#P45 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Expression of olfactory signaling molecules in the nonchemosensory<br />

tissues<br />

Nana Kang 1 , Hyoseon Kim 1 , Frank Margolis 2 , JaeHyung Koo 1<br />

1<br />

DGIST/Department of Brain Science Daegu, South Korea,<br />

2<br />

University of Maryland/Department of Anatomy and Neurobiology<br />

Baltimore, MD, USA<br />

Olfactory sense is mediated by specialized olfactory receptor<br />

neurons (ORNs) in the nose. However, ectopic expressions<br />

and functional roles of olfactory receptors (ORs) and olfactory<br />

signaling molecules (OMP, Ga olf<br />

, and AC3) still remain to be<br />

elucidated. This study demonstrates the presence of olfactory<br />

signaling molecules in non-olfactory tissues by systematically<br />

using RT-PCR, western blotting, immunohistochemistry, and<br />

a double-antibody immunoprecipitation/immunodetection<br />

procedure. Unexpectedly, the co-localization of OMP/AC3/<br />

Ga olf<br />

was confirmed in several tissues while they were expressed<br />

on different cell types of the same organ in another nonchemosensory<br />

tissue. Additionally, gene expression of olfactory<br />

receptors (ORs) was observed in non-olfactory tissues through<br />

RT-PCR. These results suggest that olfactory receptors play<br />

an important role in tissue-specific or common physiological<br />

functions of ectopic expression in non-olfactory tissues. In the<br />

future, we need to define the physiological function of olfactory<br />

receptors in non-chemosensory tissues. Acknowledgements:<br />

DGIST MIREBrain and Convergence Science Center<br />

(13-BD-0403)<br />

#P46 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

In vivo dynamic interactions between the methyl-CpG binding<br />

protein MeCP2 and chromatin under odor-evoked activity<br />

Wooje Lee, Qizhi Gong<br />

University of Cali<strong>for</strong>nia at Davis/Cell Biology and Human Anatomy<br />

Davis, CA, USA<br />

MeCP2 was identified as a methyl-CpG binding protein<br />

and capable of recruiting co-repressor complexes to<br />

promoters to suppress gene expression. MeCP2 is abundant<br />

in neurons. Mutations in MECP2 cause Rett syndrome, a<br />

neurodevelopmental disorder. Recent studies suggest that Mecp2<br />

has multiple functions including transcriptional repression/<br />

activation and structural compaction of chromatin. Dynamic<br />

interaction between MeCP2 and chromatin is not well<br />

understood. The complexity of MeCP2 function among different<br />

neuronal populations and a different methylation status in in<br />

vitro culture system have made it challenging to understand<br />

MeCP2 binding profile and dynamics under neuronal activity<br />

in vivo. Olfactory epithelium provides an ideal in vivo model<br />

in its ubiquitous neuronal population and accessibility <strong>for</strong><br />

neuronal activity manipulation. In this study, we sought to<br />

identify MeCP2 binding profiles to different regions of the<br />

chromosome and changes under odor-evoked activity. Chromatin<br />

immunoprecipitation following high throughput sequencing<br />

shows MeCP2 binds to not only methylated CpG island but<br />

also intergenic and intronic regions and sparsely methylated<br />

promoters. Genome-wide profiling <strong>for</strong> MeCP2 binding in<br />

vivo clearly shows two distinct distributions of MeCP2, one<br />

concentrated at regulatory regions and the other along the<br />

entire genes locus. Odor-evoked activity results in significant<br />

changes in MeCP2 affinity to selected gene loci. Comparing<br />

methylation state and MeCP2 binding profiles revealed that<br />

odor-evoked activity alters MeCP2 affinity to chromatin in<br />

a DNA methylation independent manner. Our results reveal<br />

the complexity of MeCP2 and chromatin interaction. We<br />

hypothesize that Mecp2 regulates activity-dependent gene<br />

regulations via changing its binding affinity to the entire gene<br />

locus. Acknowledgements: NIH DC11346<br />

#P47 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Sensory inputs modulate olfactory cilia morphology and<br />

function in the mammalian nose<br />

Rosemary Lewis 1 , Huikai Tian 1 , Jiwei He 1 , Jianbo Jiang 2 , Timothy<br />

Connelly 1 , Kai Zhao 2 , Minghong Ma 1<br />

1<br />

Department of Neuroscience, Perelman School of Medicine at the<br />

University of Pennsylvania Philadelphia, PA, USA, 2 Monell Chemical<br />

Senses Center Philadelphia, PA, USA<br />

By converting environmental signals into intracellular responses,<br />

cilia are critical <strong>for</strong> many biological processes, including<br />

olfaction. Surprisingly, little is known about what factors shape<br />

cilia morphology and how morphology impacts function.<br />

We recently discovered that olfactory cilia vary considerably<br />

in length depending on the cell location within the olfactory<br />

epithelium. Using specific markers <strong>for</strong> a subset of olfactory<br />

sensory neurons (OSNs) from C57BL/6 mice (3-6 weeks, n = 19<br />

animals), we found that cilia length increases from ~1 μm in the<br />

posterior nasal septum to ~20 μm in the anterior septum, with<br />

the longest cilia (up to 50 μm) typically found in the dorsal recess.<br />

We then built a 3D computational fluid dynamics model based<br />

on the mouse nasal cavity and demonstrated that cilia length is<br />

positively correlated with sensory inputs, particularly odorant<br />

absorption. To determine whether sensory inputs themselves<br />

account <strong>for</strong> the cilia length pattern, we per<strong>for</strong>med unilateral naris<br />

closure on newborn mice (n = 6 animals) and immunostained<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

46


olfactory cilia four weeks later. Remarkably, cilia length was<br />

increased in the open (overstimulated) nostril. We further found<br />

that cilia length modified OSN function, as OSNs expressing<br />

a defined odorant receptor were more sensitive to odorant<br />

stimulation when they had longer cilia (n = 7) as opposed to<br />

shorter cilia (n = 8). Together, these results suggest that sensory<br />

activity may shape olfactory cilia length and, consequently, OSN<br />

function. This discovery offers novel insight into the organization<br />

and function of OSNs and into cilia biology. Acknowledgements:<br />

Supported by NIDCD grants R01DC011554 and R01DC006213.<br />

#P48 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

SUMOylation regulates the ciliary localization of olfactory<br />

signaling proteins<br />

Jeremy C McIntyre, Jeffrey R Martens<br />

University of Michigan, Department of Pharmacology Ann Arbor,<br />

MI, USA<br />

In olfactory sensory neurons (OSNs), the protein components <strong>for</strong><br />

odor detection are highly enriched in cilia, however the precise<br />

mechanisms <strong>for</strong> this localization remain poorly defined. The<br />

mechanisms <strong>for</strong> selective cilia entry may be analogous to nuclear<br />

import utilizing importin ß2, a Ran gradient and nucleoporins.<br />

A unique post-translational modification process involved in<br />

nuclear-cytosolic transport is the reversible conjugation of Small<br />

Ubiquitin-like Modifier (SUMO) proteins or SUMOylation.<br />

Bioin<strong>for</strong>matic examination reveals that both adenylate cyclase 3<br />

(AC3) and the calcium-activated chloride channel, annoctamin2<br />

(ANO2) harbor conserved SUMOylation motifs. There<strong>for</strong>e, we<br />

hypothesized that SUMOylation regulates ciliary localization of<br />

AC3 and ANO2. Coexpression of SENP2, a SUMO protease,<br />

with either AC3:GFP or ANO2:GFP in MDCKII cells prevented<br />

their normal ciliary localization. Site directed mutagenesis of the<br />

predicted SUMOylation sites also blocked ciliary localization<br />

of both proteins. To test if SUMOylation is necessary <strong>for</strong><br />

trafficking of signaling proteins in vivo, mice were dually infected<br />

with wildtype or mutant adenovirus constructs along with<br />

Arl13b:mCherry (a cilia marker). Live, en face imaging of OSNs<br />

showed wildtype AC3:GFP co-localized with Ar1l3b:mCherry<br />

in olfactory cilia, while the mutant <strong>for</strong>m failed to enter<br />

olfactory cilia. Surprisingly however, both wildtype and mutant<br />

ANO2:GFP trafficked into olfactory cilia, indicating potential<br />

other mechanisms permitting trafficking in vivo. In addition, the<br />

generation of SUMOylation sites in the related channel, ANO1<br />

was not sufficient <strong>for</strong> ciliary entry. Together our data demonstrate<br />

that SUMOylation of some signaling proteins is necessary, but<br />

not sufficient <strong>for</strong> ciliary localization. Acknowledgements: This<br />

work was supported by NIDCD grants 1R01DC009606-01<br />

(JRM) and 1F32DC011990-01 (JCM)<br />

#P49 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Functional characterization of alternative signal transduction<br />

pathways in olfactory receptor neurons<br />

Paul Scholz, Sabrina Baumgart, Katharina Klasen, Benjamin Kalbe,<br />

Hanns Hatt<br />

Ruhr-University-Bochum Bochum, Germany<br />

It is generally agreed that in olfactory sensory neurons (OSNs)<br />

the binding of odorant molecules to their specific olfactory<br />

receptor (OR) triggers a cAMP-dependent signaling cascade<br />

activating cyclic-nucleotide gated (CNG) channels. However,<br />

considerable controversy dating back more than 10 years has<br />

surrounded the question of whether phosphoinositide (PI)<br />

signaling plays a role in mammalian olfactory transduction.<br />

Early studies of PI signaling in olfaction focused solely on<br />

the classical phospholipase C (PLC) dependent pathway,<br />

demonstrating that odorants can elevate inositol triphosphate<br />

(IP3). In addition, our recent study proved that odorants<br />

stimulate both, PLC and phosphatidylinositol 3-kinases (PI3Ks)<br />

in the dendritic knobs and in olfactory cilia of rodent OSNs.<br />

In this project, we aim at characterizing the dual pathway of<br />

olfactory signaling in more detail. For this purpose, we will<br />

analyze the distribution of PI signaling upon specific odor<br />

stimulation in living OSNs via translocation imaging. The<br />

use of mOR-EG-GFP transgenic mice will allow <strong>for</strong> specific<br />

analysis of OSNs expressing the well-characterized olfactory<br />

receptor mOR-EG, which can be activated by different odorants.<br />

To investigate PI signaling in OSNs, we will use adenoviral<br />

vectors carrying two different fluorescently tagged proteins,<br />

the pleckstrin homology (PH) domains of phospholipase C<br />

(PLC) and the general receptor of phosphoinositides (GRP1),<br />

to monitor PI activity in the murine olfactory epithelium in vivo.<br />

Furthermore, we will monitor the effects of PI signaling on the<br />

electrophysiological output of OSNs by patch clamp technique<br />

of single neurons in acute OE slices of mOR-EG-GFP<br />

transgenic mice.<br />

#P50 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Phospholipase C Mediates Intracellular Ca 2+ Increase via<br />

Internal Ca 2+ Stores and trpM5 Activation in Mouse<br />

Olfactory Sensory Neurons<br />

Steven A. Szebenyi, Tatsuya Ogura, Jen Chang, Aaron Sathyanesan,<br />

Weihong Lin<br />

Department of Biological <strong>Sciences</strong>, University of Maryland,<br />

Baltimore County, Baltimore, MD, USA<br />

Phospholipase C (PLC) and internal Ca 2+ stores are involved in<br />

a variety of cellular signaling including sensory transduction.<br />

However, our understanding of the PLC pathway in mammalian<br />

olfactory sensory neurons (OSNs) is largely limited to those<br />

studies in which the PLC inhibitor U73122 was used to suppress<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

47


odor responses. Recently, transient receptor potential channel M5<br />

(trpM5) has been shown to express in a population of mature<br />

OSNs in mice (Lin et al 2007). trpM5 is an essential downstream<br />

effector of the PLC pathway <strong>for</strong> taste transduction. Here we<br />

investigate PLC, trpM5 and internal Ca 2+ stores in freshly<br />

isolated mouse OSNs using single cell Ca 2+ imaging.<br />

We found that OSNs responded to a PLC activator m-3M3FBS<br />

in a concentration dependent manner with a higher percentage<br />

of responding cells and greater amplitudes after an increase in<br />

concentration (78%, n=23 at 15µM, 90.3%, n=52 at 25µM).<br />

In contrast, only one out of 9 OSNs responded to the inactive<br />

analog o-3M3FBS (25µM). Eliminating extracellular Ca 2+ did<br />

not reduce the percent of responding OSNs to m-3M3FBS and<br />

only the response amplitudes were moderately reduced (n=7).<br />

In addition, The PLC inhibitor U73122 (5-10µM) greatly reduced<br />

the percent of OSNs responding to m-3M3FBS (37.5%, n=8)<br />

and the response amplitudes. Further, using OSNs isolated<br />

from trpM5-GFP and trpM5 knockout-GFP mice, we found<br />

that trpM5-expressing OSNs responded to m-3M3FBS with a<br />

significantly larger amplitude and calcium load than the trpM5-<br />

null OSNs (n= 6 to 9 <strong>for</strong> each group). Our data suggest that<br />

most OSNs are capable of utilizing the PLC pathway to release<br />

Ca 2+ from internal Ca 2+ stores and that subsequent activation of<br />

trpM5 in trpM5-expressing OSNs leads to additional increases<br />

in intracellular Ca 2+ loads. Acknowledgements: Supported by<br />

research grants NIH/NIDCD 009269, 012831 and ARRA<br />

administrative supplement to WL<br />

#P51 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

mouse model to understand the functional role of Q8BH53 in<br />

the olfactory system. In addition, we are taking a biochemical<br />

approach to identify binding partners of Q8BH53 in OSN cilia.<br />

Acknowledgements: NIH DC007395<br />

#P52 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Expression of Several Odorant Receptors Outside the<br />

Olfactory System<br />

Crystal M. Wall, Marsalis Brown, Haiqing Zhao<br />

Johns Hopkins University/Biology Baltimore, MD, USA<br />

Over one thousand G-protein coupled receptors have been<br />

classified as odorant receptors in the mouse genome based<br />

on sequence similarities with receptors found in the olfactory<br />

system. Recent studies have identified odorant receptors<br />

expressed outside of the olfactory system, and growing evidence<br />

supports a role <strong>for</strong> these receptors in extra-olfactory functions.<br />

We have examined expression of the odorant receptors M71,<br />

M72, I7, P2, MOR28, whose expression patterns have been<br />

well characterized in the olfactory system. We screened <strong>for</strong><br />

expression of these receptors in extra-olfactory tissues by<br />

RT-PCR and then tested positive results by in situ hybridization,<br />

immunohistochemistry, or through the use of established<br />

reporter mouse lines. Identification of novel expression outside<br />

the olfactory system suggests a function <strong>for</strong> odorant receptors in<br />

these tissues. Future research will be aimed toward uncovering<br />

functions and identifying putative endogenous ligands.<br />

Acknowledgements: NIH DC007395<br />

Analysis of the novel protein Q8BH53 in olfactory<br />

sensory neuron cilia<br />

Anna K Talaga 1 , Aaron B Stephan 2 , Varun Chokshi 1 , Haiqing Zhao 1<br />

1<br />

Johns Hopkins University/Biology Baltimore, MD, USA,<br />

2<br />

UCSD/Division of Biological <strong>Sciences</strong> La Jolla, CA, USA<br />

The cilia of olfactory sensory neurons (OSNs) are specialized<br />

<strong>for</strong> encoding and transducing odor in<strong>for</strong>mation. Among the<br />

proteins found in the cilia, many are critical <strong>for</strong> mediating<br />

and/or modulating olfactory signal transduction. We detected<br />

a novel protein, Q8BH53, from a proteomic screen of OSN<br />

cilial membrane preparations. Q8BH53 is conserved among<br />

eukaryotes and is a unique protein, as no other paralogs exist<br />

in the mouse genome. Bioin<strong>for</strong>matic analysis suggested that<br />

the majority of the Q8BH53 sequence is composed of ARMdomains.<br />

Although the function of Q8BH53 is unknown, the<br />

presence of these ARM-repeat domains signifies that it may<br />

be important <strong>for</strong> establishing protein-protein interactions.<br />

q8bh53 transcripts are abundant in the mouse olfactory<br />

epithelium and are also found in several other tissues. Using<br />

immunohistochemistry, we found that Q8BH53 localizes<br />

specifically to OSN cilia but is largely excluded from the<br />

respiratory epithelium cilia in adult mice. Furthermore, we found<br />

that Q8BH53 expression begins around embryonic day E13.5<br />

in the olfactory epithelium. We are currently using a knock-out<br />

#P53 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

M3-R inhibits b-arrestin2 recruitment and desensitization of<br />

mammalian odorant receptors to potentiate their activation<br />

Yue Jiang 1 , Yun R. Li 2 , Hiroaki Matsunami 1,3<br />

1<br />

Duke University Durham, NC, USA, 2 University of Pennsylvania,<br />

Perelman School of Medicine Philadelphia, PA, USA, 3 Duke University<br />

Medical Center Durham, NC, USA<br />

Adjusting sensitivity of sensory stimuli needed by an animal at<br />

any given time is crucial <strong>for</strong> animals’ survival. In mammals, the<br />

activation of odorant receptors (ORs), which are expressed in the<br />

olfactory sensory neurons (OSNs) in the olfactory epithelium,<br />

mediates the perception of smell. These ORs belong to the<br />

large family of G protein-coupled receptors (GPCRs), which<br />

also include the muscarinic acetylcholine receptors that are<br />

key mediators of the parasympathetic nervous system output.<br />

Previously, we showed that activation of the muscarinic receptor<br />

M3-R has been shown to potentiate OR-mediated cAMP<br />

response, and the functional interaction between the M3-R<br />

and ORs suggests that odorant detection may be modulated<br />

by the neurotransmitter at the peripheral level. However, the<br />

mechanisms underlying this modulation were not understood.<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

48


Here we provide evidence suggesting that M3-R mediates the<br />

potentiation of OR signaling by inhibiting the recruitment of<br />

b-arrestin-2 to activated ORs. In line with this, activation of<br />

the M3-R by muscarinic agonist further inhibits b-arrestin-2<br />

recruitment to ORs, while M3-R antagonist alleviates the<br />

inhibition. These effects are not explained by the competition <strong>for</strong><br />

b-arrestin-2 between the two receptors. Further more, the third<br />

intracellular loop of the M3-R is responsible <strong>for</strong> its regulation of<br />

OR activity. This data suggests that the M3-R potentiates ORmediated<br />

cAMP response largely by inhibiting the b-arrestin-2<br />

recruitment to ORs, providing evidence <strong>for</strong> a novel mechanism<br />

of OR activity regulation by non-OR GPCRs, with b-arrestin-2<br />

as a crucial mediator. Acknowledgements: This work is<br />

supported by grant from NIH.<br />

#P54 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Off-flavor substances in foods and beverages cause a potent<br />

suppression of olfactory signal transduction<br />

Hiroko Takeuchi 1 , Hiroyuki Kato 2 , Takashi Kurahashi 1<br />

1<br />

Osaka University Osaka, Japan, 2 Daiwa Can Company Tokyo, Japan<br />

We examined the effect of off-flavors in foods/beverages<br />

on the olfactory receptor cells under the whole-cell voltage<br />

clamp recording configuration. Generally, it has been shown<br />

that off-flavor substances induce exogenous unpleasant smells<br />

even with a very low concentration (ppt level inclusion in<br />

products). Although not yet scientifically demonstrated, it has<br />

also been pointed out that off-flavors reduce the pleasant flavors<br />

contained in foods/beverages. One of the most powerful offflavors<br />

is 2,4,6-trichloroanisole (TCA), that is especially known<br />

<strong>for</strong> inducing the corktaint of wines. In the present study, we<br />

show with human psychophysical tests that TCA and related<br />

substances actually reduce flavors of foods and beverage with<br />

very low concentration. It was shown that TCA also suppressed<br />

cyclic nucleotide-gated (CNG) channels potently, when<br />

examined in olfactory sensory cilia. Surprisingly, the channel<br />

suppression was detected even when 1 aM of TCA was applied<br />

to the cell with an U-tube system. To explain such superefficiency,<br />

the TCA effect showed the time-integration and slow<br />

recovery from the current suppression, presumably representing<br />

the integration of the substance into the hydrophobic site of the<br />

membrane. Based on the relation between the number of TCA<br />

molecules applied and total number of CNG channels, it was<br />

assumed that single TCA molecule may affect more than one<br />

CNG channel. This is also consistent with an idea that the<br />

effect of TCA is mediated by the lipid bilayer to affect<br />

surrounding channels simultaneously. TCA was found in a<br />

wide variety of foods/beverages, when screened out from their<br />

flavor losses. Natural generation of TCA and related off-flavor<br />

substances may be one of the mechanisms <strong>for</strong> the deterioration<br />

of those products.<br />

#P55 POSTER SESSION I:<br />

MULTIMODAL RECEPTION; CHEMOSENSATION<br />

AND DISEASE; OLFACTION PERIPHERY<br />

Effect of Vitamin A Deficiency on Olfactory Marker Protein<br />

Expression in Postnatal Mouse Olfactory Neurons<br />

Bukola A. Oke, Mary Ann Asson-Batres<br />

Tenneessee State University Biology Department Nashville, TN, USA<br />

Our lab has previously shown that Olfactory receptor neuron<br />

(ORN) levels decrease when sources of Vitamin A have been<br />

removed from the diet of postnatal rats. Experiments per<strong>for</strong>med<br />

have been directed towards discovering whether or not the<br />

same results are observed in mice that were mutated with a<br />

lecithin retinyl acyl transferase (LRAT knockout, KO) gene.<br />

To determine the VAD effects on mice, specifically on ORN<br />

numbers we took LRAT KO male mice who were age-matched<br />

and fed them a VAD diet or a diet supplemented with VA (VAS)<br />

<strong>for</strong> 8 weeks (Study 1) or 19 weeks (Study 2). Studies included our<br />

control group, age matched, male wild type (WT) VAS mice. Our<br />

VAD rats exhibit distinct signs of VAD that include alopecia,<br />

ataxia, reddened eyes, white teeth and decreased weight gain<br />

relative to controls, LRAT KO mice did not display any of these<br />

signs after eight weeks on the VAD diet. LRAT KO mice on the<br />

VAD diet <strong>for</strong> 19 weeks showed a relative weight loss in the latter<br />

part of the study. Cytosolic extracts from Olfactory tissue were<br />

collected from mice in Study 2. Immunoblots were prepared<br />

and probed with an antibody directed against olfactory marker<br />

protein (OMP), a marker <strong>for</strong> mature ORNs. Chemiluminescent<br />

reagent detection system and images were acquired with a<br />

BioImaging System were used to evaluate relative OMP protein<br />

expression levels in LRAT KO VAD, LRAT KO VAS, and WT<br />

mice olfactory tissue cytosolic lysates. Signal area densities<br />

were recorded using an EpiChemi Darkroom UVP Bioimaging<br />

System. This allowed us to quantify the density of each band on<br />

the immunoblot. Results have shown that VAD mice have less<br />

OMP expression than those animals fed a VAS diet including the<br />

LRAT KO VAS mice and the WT mice. These findings suggest<br />

that VAD may have different and/or less pronounced effects<br />

on mice than rats. Acknowledgements: NIH/NIGMS/MBRS/<br />

SCORE S06 GM 008092<br />

#P56 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Cerebral processing of odors related to their hedonic judgment<br />

Anna Blumrich, Cornelia Hummel, Emilia Iannilli, Thomas Hummel<br />

Smell & Taste Clinic, Department of Otorhinolaryngology, University<br />

of Dresden Medical School Dresden, Germany<br />

The aim of the study was to investigate parameters reflecting<br />

the hedonic component of odor perception. The emotional<br />

effect of smelling- claiming an odor as pleasant or unpleasantis<br />

an important part of the central nervous connection of odor<br />

perception. It has been shown that perception of pleasant odors<br />

implicates different cerebral activations than that of unpleasant<br />

ones. Thirty-two healthy, right-handed subjects (16 men, 16<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

49


women; mean age 23.4 years; range 20-29 years) were examined.<br />

Normal olfactory function was ascertained using the “Sniffin<br />

Sticks”. All subjects rated peach odor as pleasant, and the smell<br />

of butanol as unpleasant. According to their hedonic judgement<br />

of liquorice odor, subjects were divided into two groups: group<br />

1 subjects described the odor of liquorice as pleasant, and group<br />

2 subjects experienced liquorice as unpleasant. Functional<br />

magnetic resonance imaging (fMRI) was used to compare the<br />

cerebral activations while smelling the three different odors:<br />

pleasant peach, unpleasant butanol, and ambiguous liquorice.<br />

Analysis indicated common neural activations in response to<br />

all odors in a number of regions, e.g. in cingulate gyrus, medial<br />

frontal gyrus and caudate nucleus. In subjects disliking liquorice,<br />

activations were found in four areas including medial frontal<br />

gyrus, postcentral gyrus and medial temporal gyrus. In liquoriceliking<br />

subjects, activations were found in the medial frontal<br />

lobe. These results suggest that there are different patterns of<br />

central nervous activation dependent on the hedonic value of an<br />

odor, mainly including the frontal lobe, with unpleasant odors<br />

activating the more lateral frontal lobe and pleasant odors more<br />

medial areas. In addition, unpleasant odors produced more<br />

activation as compared to pleasant odors. Acknowledgements:<br />

Supported by Takasago, Paris.<br />

#P57 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Valence Modulation of Crossmodal Olfactory-Visual<br />

Neural Integration<br />

Jessica Freiherr 1 , Anna-Nora zur Nieden 1 2, 3, 4<br />

, Johan N. Lundström<br />

1<br />

RWTH Aachen University, Diagnostic and Interventional<br />

Neuroradiology Aachen, Germany, 2 Monell Chemical Senses Center<br />

Philadelphia, PA, USA, 3 University of Pennsylvania, Department of<br />

Psychology Philadelphia, PA, USA, 4 Karolinska Institute, Department<br />

of Clinical Neuroscience Stockholm, Sweden<br />

We recently demonstrated that a concurring congruent visual<br />

stimulus does not affect olfactory sensitivity but does modulate<br />

the perceived valence or pleasantness and intensity of an odor.<br />

However, the congruency-dependent effect occurred only <strong>for</strong><br />

odors perceived as pleasant. With the current study we explored<br />

the neural mechanisms of this behavioral phenomenon with<br />

the aim of determining the influence of valence on the neural<br />

correlates of crossmodal olfactory-visual integration. To this<br />

end, the pleasant odor phenyl ethyl alcohol and unpleasant odor<br />

isovaleric acid was applied using constant-flow olfactometry<br />

in combination with a congruent, incongruent, or blank visual<br />

stimulus during an event-related fMRI paradigm. As control<br />

stimuli we also applied pleasant and unpleasant visual stimuli<br />

and a baseline stimulus. We investigated brain activation due to<br />

crossmodal integration in 14 healthy, normosmic participants.<br />

Subjects had to rate pleasantness of the odors after each<br />

event. Statistical analyses of the behavioral data demonstrate<br />

a replication of the a<strong>for</strong>ementioned findings. As predicted,<br />

valence-independent olfactory-visual integration was mediated<br />

by low-level multisensory integration areas in the superior<br />

parietal lobule. Preliminary analyses of the fMRI data indicate<br />

that valence-dependent integration occurs in higher-order<br />

multisensory integration areas in conjunction with areas known<br />

to code <strong>for</strong> odor valence. Moreover, a differential processing<br />

of unpleasant compared to pleasant olfactory-visual stimulus<br />

combinations has been established. Further insights into the<br />

neural processes mediating the influence of valence on the<br />

neural correlates of olfactory-visual integration will be discussed.<br />

Acknowledgements: Funded by a startup grant from the Medical<br />

Faculty of RWTH Aachen University.<br />

#P58 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Visuo-olfactory integration facilitates peri-threshold<br />

olfactory categorization<br />

Jaryd Hiser 1 , Lucas R. Novak 1 , Wen Li 1,2<br />

1<br />

Department of Psychology, University of Wisconsin - Madison<br />

Madison, WI, USA, 2 Waisman Center, University of Wisconsin -<br />

Madison Madison, WI, USA<br />

Olfactory quality discrimination and categorization proves<br />

a highly challenging process in humans. Additional in<strong>for</strong>mation<br />

from other senses, such as visual cues, may facilitate this<br />

operation via crossmodal integration. To date, crossmodal<br />

sensory integration has focused on non-chemical senses. Using<br />

functional magnetic resonance imaging (fMRI) techniques,<br />

this study characterized visuo-olfactory integration in olfactory<br />

categorization. Participants (N=29) smelled an odor at a merely<br />

detectable level from one of three categories (food, floral, or<br />

wood) while viewing a picture that was congruent or incongruent<br />

to the odor, and then made a category decision on the odor.<br />

Reaction time (RT) was faster <strong>for</strong> congruent versus incongruent<br />

stimuli (P


#P59 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

#P60 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

The chemosensory path of pain<br />

Kathrin Kollndorfer 1 , Ksenia Kowalczyk 1 , Johannes Frasnelli 2 ,<br />

Elisabeth Hoche 1 , Ewald Unger 3 , Christian A. Mueller 4 ,<br />

Siegfried Trattnig 5 , Veronika Schöpf 1<br />

1<br />

Medical University of Vienna, Department of Radiology, Division of<br />

Neuro- and Musculoskeletal Radiology Vienna, Austria, 2 CERNEC,<br />

Université de Montréal, Département de Psychologie Montreal, QC,<br />

Canada, 3 Medical University of Vienna, Center <strong>for</strong> Medical Physics<br />

and Biomedical Engineering Vienna, Austria, 4 Medical University of<br />

Vienna, Department of Otorhinolaryngology Vienna, Austria, 5 Medical<br />

University of Vienna, Department of Radiology, MR Centre of<br />

Ecxellence Vienna, Austria<br />

The function of the olfactory system has been well investigated<br />

over the last decades. However, we know much less regarding<br />

central processing of intranasal trigeminal stimuli. Trigeminal<br />

sensations like burning, stinging, tingling, pungency and<br />

temperature constitute an additional quality in perception of<br />

odor and flavor, but also allow odor localization in contrast<br />

to pure olfactory odors. In this study we focused on three<br />

different trigeminal compounds which target various sensory<br />

receptors and produce distinctive sensations: CO 2<br />

elicits stinging<br />

sensations (activating TRPV1 receptors), cinnamaldehyde<br />

burning sensations (TRPA1) and menthol cooling sensations<br />

(TRPM8). The point of interest was to investigate discrepancies<br />

in the processing of different intranasal chemosensory trigeminal<br />

stimulations along the trigeminal pathway. Moreover, we aimed<br />

to reveal activation in olfactory areas which supports a close<br />

relationship between the two chemosensory systems.<br />

We conducted an fMRI study on a 3T MRI scanner to measure<br />

neural activity in 12 healthy subjects. Each of the three stimuli<br />

was delivered separately by a computer-controlled air-dilution<br />

olfactometer in one session. Using an event-related design<br />

subjects received trigeminal stimuli of 500ms to the left nostril<br />

(ISI 30s). During the whole experiment subjects were asked to<br />

breathe using velopharyngeal closure. Functional imaging data<br />

were analyzed with Independent Component Analysis. Insula,<br />

SI and OFC were activated by all three compounds. Further, all<br />

three conditions led to activation of the PFC demonstrating the<br />

well established interaction between the olfactory and trigeminal<br />

system. The findings support the notion that activation of the<br />

three trigeminal receptors is processed in the same pathway.<br />

Acknowledgements: FWF (P23205-B09)<br />

Smell what you seek – Reward sensitivity amplifies<br />

visuo-olfactory integration of positive affect<br />

Lucas R. Novak 1 , Jaryd Hiser 1 , Wen Li 1,2<br />

1<br />

Department of Psychology, University of Wisconsin - Madison<br />

Madison, WI, USA, 2 Waisman Center, University of Wisconsin -<br />

Madison Madison, WI, USA<br />

Crossmodal integration is a process ubiquitous in animals with<br />

multiple sensory systems, facilitating perception especially<br />

when challenged with limited stimulus input. However, there<br />

is little research on integration of positive affect, still less how<br />

this process varies with relevant individual differences such as<br />

reward sensitivity. Reward sensitivity may increase response to<br />

rewarding stimuli, thereby facilitating crossmodal integration<br />

of positive affect. Applying functional magnetic resonance<br />

imaging (fMRI) techniques, this study examined visual and<br />

olfactory integration of reward. Participants (N=29) per<strong>for</strong>med<br />

an odor categorization task: they smelled an odor from one<br />

of two categories (floral or wood) while viewing an image<br />

congruent or incongruent with the odor, followed by a category<br />

decision. Reaction time was faster and accuracy greater <strong>for</strong><br />

floral than wood odors (P’s


#P61 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

#P62 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Biochemical components of trigeminal integration in<br />

anosmics: A pilot functional magnetic spectroscopy study<br />

Veronika Schöpf 1 , Kathrin Kollndorfer 1 , Elisabeth Hoche 1 , Bernhard<br />

Strasser 2 , Ksenia Kowalczyk 1 , Ewald Unger 3 , Christian A. Müller 4 ,<br />

Siegfried Trattnig 2 , Martin Krssak 2,5<br />

1<br />

Department of Radiology, Division of Neuro- and Musculoskeletal<br />

Radiology Medical University of Vienna Vienna, Austria,<br />

2<br />

Department of Radiology, MR Centre of Excellence, Medical<br />

University of Vienna Vienna, Austria, 3 Center <strong>for</strong> Medical Physics<br />

and Biomedical Engineering, Medical University of Vienna Vienna,<br />

Austria, 4 Department of Otorhinolaryngology, Medical University<br />

of Vienna Vienna, Austria, 5 Department of Internal Medicine III,<br />

Division of Endocrinology and Metabolism, Medial University of<br />

Vienna Vienna, Austria<br />

Several studies investigated functional interaction of the olfactory<br />

and trigeminal system. Although anosmic patients are not able to<br />

perceive odors they show changes in “insular” blood oxygenation<br />

during trigeminal neuronal activity. This pilot study aimed to<br />

investigate how changes on blood oxygenation level, as observed<br />

by fMRI, are underlined by changes in neurotransmitter (GABA,<br />

glutamate) balance under trigeminal simulation in the insula<br />

and how these changes differ between anosmic and healthy<br />

populations. 3 (2f;25-43ys) functional anosmics and 8 controls<br />

(6f;18-43ys) were examined using proton magnetic resonance<br />

spectroscopy (1H-MRS) to explore changes of excitatory and<br />

inhibitory neurotransmitters in the insula. Spin-echo based MRS<br />

(TE=30ms/TR=5000ms) measurements were per<strong>for</strong>med on a<br />

3T whole body MR scanner, using a 32ch coil <strong>for</strong> signal<br />

detection combined with a stimulation device, which was<br />

designed specifically <strong>for</strong> intranasal application. The paradigm<br />

consisted of 4 stimulation blocks: 4 dynamic cycles with 32<br />

acquisitions, and 32 stimuli (CO 2<br />

, 50%v/v, birhinal, 250ms).<br />

Acquired spectral transients were individually frequency<br />

corrected, phased and further grouped according to the<br />

timeline position into baseline- and stimulation-groups. The<br />

quantification of metabolic intensities was per<strong>for</strong>med using<br />

the LCModel with an imported modelled basis set including<br />

metabolites and macromolecular resonances. Results <strong>for</strong> controls<br />

revealed decreased GABA (41.24%) during the rest phase.<br />

Anosmic patients showed a significant decrease of glutamate<br />

(16.42%) and a higher GABA response (205.25%) rate to<br />

stimulation compared to controls. Results of this study will<br />

significantly contribute to the basic understanding of trigeminal<br />

processing of chemosensory in<strong>for</strong>mation in patients with<br />

olfactory dysfunction. Acknowledgements: FWF (P23205-B09)<br />

Superadditive processing during flavor perception is<br />

modulated by anterior temporal cortex connectivity<br />

Janina Seubert 1 , Kathrin Ohla 1,2 , Yoshiko Yokomukai 3 ,<br />

Johan N. Lundström 1,4,5<br />

1<br />

Monell Chemical Senses Center Philadelphia, PA, USA,<br />

2<br />

German Institute of Human Nutrition Potsdam-Rehbrücke,<br />

Germany, 3 Kirin Holdings Company LTD Tokyo, Japan,<br />

4<br />

Department of Psychology, University of Pennsylvania Philadelphia,<br />

PA, USA, 5 Department of Clinical Neurosience, Karolinska Institute<br />

Stockholm, Sweden<br />

The combination of taste and odor found in a flavorful dish<br />

creates a more powerful sensation than its odor or taste in<br />

isolation. Whereas the neural processing of the individual<br />

chemosensory components is well known, the functional<br />

connectivity underlying the combined flavor percept is poorly<br />

understood. In the present functional magnetic resonance<br />

imaging study, subjects were presented with taste only<br />

(gustatory presentation of juice, closed soft palate), smell only<br />

(orthonasal presentation of juice odor), or a combined flavor<br />

(retronasal-gustatory presentation, swallowing juice). As<br />

expected, olfactory stimulation alone activated olfactory areas<br />

while gustatory stimulation alone elicited activation within the<br />

gustatory cortex. Overlapping activation within both networks<br />

could be observed during flavor presentation, and a convergence<br />

zone between all three conditions was observed in the anterior<br />

ventral insula and cingulate cortex. Superadditive activity <strong>for</strong> the<br />

flavor condition, relative to odor and taste alone, was observed<br />

in the dorsal insular gyrus, extending into parietal operculum<br />

and postcentral gyrus. Finally, to delineate the cerebral networks<br />

contributing to the flavor percept, we assessed the functional<br />

connectivity between these significant nodes responsive to<br />

chemosensory overlap during combined odor-taste stimulation.<br />

Increases in functional connectivity with both convergent and<br />

superadditive areas were observed in an overlapping area in<br />

the temporal pole. Taken together, these findings are suggestive<br />

of an important relay function of semantic memory circuits<br />

in the <strong>for</strong>mation of the flavor experience from crossmodal<br />

chemosensory in<strong>for</strong>mation.<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

52


#P63 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

#P64 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Is the Superior Temporal Sulcus involved in the Rein<strong>for</strong>ced<br />

Configural Processing of a Binary Odor Mixture?<br />

Charlotte Sinding 1,2 , Gérard Coureaud 1 , Noëlle Béno 1 , Elodie Le Berre 1 ,<br />

Cornelia Hummel 2 , John Prescott 3 , Moustafa Bensafi 4 ,<br />

Thomas Hummel 2 , Thierry Thomas-Danguin 1<br />

1<br />

Centre des <strong>Sciences</strong> du Goût et de l’Alimentation (CSGA), UMR 6265<br />

CNRS, UMR 1324 INRA, Université de Bourgogne, Developmental<br />

Ethology and Cognitive Psychology Team and Flavour Perception<br />

Team Dijon, France, 2 Smell & Taste Clinic, Department of<br />

Otorhinolaryngology, University of Dresden Medical School Dresden,<br />

Germany, 3 TasteMatters Research & Consulting Sydney, Australia,<br />

4<br />

Centre de Recherche en Neurosciences de Lyon, UMR5020 CNRS,<br />

Université Lyon 1, Neurosciences Sensorielles Comportement Cognition<br />

Lyon, France<br />

In macaque monkeys the superior temporal sulcus (STS)<br />

projects to the orbitofrontal cortex which projects to the<br />

primary olfactory cortex (Carmichael and Price 1995). These<br />

connections are believed to be reciprocal, there<strong>for</strong>e the STS is<br />

supposed to receive inputs from olfactory areas. Kettenmann<br />

et al. (1996) showed that STS was activated when participants<br />

were stimulated with monomolecular odors. This area is also<br />

specifically activated <strong>for</strong> configural visual processing, as during<br />

the perception of body motion (Thompson et al. 2005). In the<br />

present study, we investigated the configural processing of odor<br />

mixtures in human adults. We repeatedly exposed (2 sessions of<br />

11 exposures) healthy volunteers (n=12, G AB<br />

) to a binary mixture<br />

(AB) configurally processed (blending of the two components’<br />

odors into a single pineapple odor), while others (n=14, G compo<br />

)<br />

were exposed to the separate components, A (“strawberry”)<br />

and B (“caramel”). To equilibrate the number of exposures in<br />

the two groups, G AB<br />

was also exposed to PEA (monomolecular,<br />

“rose”). Such exposures were known to favor the perception of<br />

the AB configuration in G AB<br />

and the perception of the elements<br />

in G compo<br />

(Sinding et al. 2011, in preparation). Two days after<br />

the pre-exposure, all subjects received an fMRI while stimulated<br />

by AB, A, B and PEA. As a major result, the STS appeared<br />

significantly more activated <strong>for</strong> the processing of the AB mixture<br />

in G AB<br />

than in G compo<br />

. The STS was also more activated <strong>for</strong> the<br />

processing of AB as compared to A and B, in G AB<br />

. The STS was<br />

not significantly activated in G compo<br />

<strong>for</strong> any stimulation, in any<br />

contrast. These results suggest that the STS is a critical area <strong>for</strong><br />

the rein<strong>for</strong>ced configural processing of simple odorant mixtures.<br />

However, PEA also activated significantly the STS in G AB<br />

in<br />

comparison to G compo<br />

. Acknowledgements: Supported by grants<br />

from the Burgundy Regional council and EU-ERDF to GC and<br />

TTD, European Dijon-Dresden Laboratory (LEA 549) to TH,<br />

GC, TTD, and a fellowship from the French MESR to CS.<br />

<strong>Association</strong> of Pleasantness and Intensity of Sweet and<br />

Salty Taste in the Human Brain<br />

Jianli Wang 1 , Sebastian Rupprecht 1 , Zachary Mosher 1 ,<br />

Robert Mchugh 1 , Jeffrey Vesek 1 , Sarah Ryan 1 , Megha Vasavada 1 ,<br />

Qing X. Yang 1,2 , Andras Hajnal 3,4 , Ann M. Rogers 4<br />

1<br />

Penn State College of Medicine/Radiology Hershey, PA, USA,<br />

2<br />

Penn State College of Medicine/Neurosurgery Hershey, PA, USA,<br />

3<br />

Penn State College of Medicine/Neural & Behavioral <strong>Sciences</strong> Hershey,<br />

PA, USA, 4 Penn State College of Medicine/Surgery Hershey, PA, USA<br />

An abnormal assessment of pleasantness and intensity of<br />

tastants can lead to negative, long-term health conditions.<br />

Among these, obesity may be a consequence of diminished<br />

sensitivity to sweet tastes, leading to amplified cravings and<br />

associated weight gains. An understanding of the neural<br />

processes of taste perception is essential <strong>for</strong> elucidating this<br />

disease state. Un<strong>for</strong>tunately, studies on the human brain are<br />

limited, and thus the majority of neural taste function knowledge<br />

stems from animal studies. These studies have shown that<br />

hedonic and intensity in<strong>for</strong>mation is encoded in the primary<br />

gustatory cortex, with neuronal firing correlated with stimulation<br />

intensity. Seeking to confirm animal models in the human<br />

brain, this study utilized fMRI to study BOLD signal changes<br />

in response to varied concentrations of sweet and salty taste<br />

stimulants. Normal, healthy volunteers (n=11) completed a<br />

total of fifteen taste fMRI studies in which brain response to an<br />

event-related taste stimulation paradigm was correlated with<br />

perceived ratings of both pleasantness and intensity. The data<br />

indicate a positive correlation between pleasantness and intensity<br />

<strong>for</strong> sweet tastants; while a negative correlation was observed <strong>for</strong><br />

salty tastants. Each triggered significant activation in the primary<br />

and secondary gustatory cortices including: bilateral anterior<br />

insular cortex, posterior orbitofrontal cortex, cingulated cortex,<br />

and dorsolateral prefrontal cortex. BOLD signals in these regions<br />

were significantly correlated with both hedonic and intensity<br />

ratings from subjects. Overall these results show that the human<br />

brain processes hedonic and intensity in<strong>for</strong>mation of sweet and<br />

salty taste through similar neural networks previously seen in<br />

animal models.<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

53


#P65 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

#P66 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Perception and encoding of odor elements and mixtures<br />

in the human brain<br />

Keng Nei Wu 1 , Sydni M. Cole 1 , Jay A. Gottfried 1,2<br />

1<br />

Northwestern University/Neurology Department, Feinberg School of<br />

Medicine Chicago, IL, USA, 2 Northwestern University/Department of<br />

Psychology Evanston, IL, USA<br />

In the natural environment, most odorous objects are composed<br />

of dozens, if not hundreds, of volatile molecules. Despite<br />

this apparent complexity, the olfactory system seamlessly<br />

integrates these components into perceptual wholes. Utilizing<br />

a between subject design, this experiment aimed to investigate<br />

how experience in the <strong>for</strong>m of aversive learning modulates<br />

perception and encoding of odor mixtures by pairing either a<br />

target binary odor mixture (Mx) or one of its components (Ele)<br />

with an electric shock. We used psychophysical measurements,<br />

functional magnetic resonance imaging (fMRI), and multivariate<br />

analytical techniques to investigate these learning induced<br />

changes. We presented human subjects with six stimuli: three<br />

monomolecular odorants (A, B, C), and three binary mixtures<br />

(AB, BC, AC). To date, results have been collected <strong>for</strong> 13 subjects<br />

(8 Ele, 5 Mx) who were successfully conditioned. When asked to<br />

identify the odor component(s) of these stimuli, subjects in the<br />

Mx group showed decreased accuracy in identifying the correct<br />

component(s). Moreover, these subjects also rated mixtures to be<br />

less similar to their components, while subjects in the Ele group<br />

rated mixtures to be more similar to their components after<br />

conditioning. These preliminary findings suggest that olfactory<br />

learning of a binary mixture may induce perceptual and neural<br />

fusion of odor elements into a synthetic whole. Conversely,<br />

pairing a shock with a component of a binary mixture may<br />

induce neural “fission” of the mixture, such that its components<br />

are processed in a more elemental fashion. Ongoing fMRI<br />

analysis will test the hypothesis that learning induced changes<br />

in odor quality perception may be reflected in the correlation<br />

between odor evoked patterns of activation in the posterior<br />

piri<strong>for</strong>m cortex. Acknowledgements: This work was supported<br />

by Northwestern Institutional Predoctoral Training Awards<br />

to K.N.W. (T32NS047987) and grants R01DC010014 and<br />

K08DC007653 from the US National Institute on Deafness and<br />

Other Communication Disorders to J.A.G.<br />

The Fate of the Inner Nose: Odor Imagery in Patients With<br />

Olfactory Loss<br />

Elena L. R. Flohr 1,2 , Artin Arshamian 3,1 , Matthias J. Wieser 2 , Cornelia<br />

Hummel 1 , Maria Larsson 3 , Andreas Muehlberger 2 , Thomas Hummel 1<br />

1<br />

Smell and Taste Clinic, University of Dresden Medical School Dresden,<br />

Germany, 2 Department of Psychology I, University of Würzburg<br />

Wuerzburg, Germany, 3 Department of Psychology, Stockholm<br />

University Stockholm, Sweden<br />

Although the concept of olfactory mental imagery remains<br />

controversial, recent studies support the principle. Cerebral<br />

activations during olfactory mental imagery are fairly well<br />

investigated in healthy participants but very few studies<br />

address the subjects of olfactory imagery in patients with<br />

olfactory loss. To investigate if olfactory imagery is impaired<br />

in patients who are no longer able to smell, 16 participants<br />

with acquired anosmia and 19 normosmic control participants<br />

have been investigated. We used functional magnet resonance<br />

tomography and subjective ratings to explore the mechanisms<br />

during mental imagery of odors. After an imagery training,<br />

participants imagined odors triggered by words naming pleasant<br />

and unpleasant olfactory objects. We found that the patients<br />

compared to healthy control participants showed greater<br />

difficulties in imagining odors and lower intensity scores while<br />

doing so. Looking at neural activation, the pattern observed<br />

by Bensafi et al. (2007) that imagining unpleasant odors leads<br />

to more activation in olfaction-related areas than imagining<br />

pleasant odors was found in the control group but not in the<br />

anosmic group. This hedonic specific approach was meant to<br />

control <strong>for</strong> activation that was due to attention allocation or<br />

activation of semantic circuits that are alone sufficient to evoke<br />

activation in olfactory areas. Direct comparisons between the<br />

groups revealed greater activation in the anosmic group in<br />

olfactory areas than in the control group. We conclude that,<br />

in contrast to the control group, anosmic participants have<br />

difficulties to per<strong>for</strong>m olfactory imagery in the conventional<br />

meaning.<br />

#P67 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Multi-modal functional imaging of rat olfactory bulb with<br />

orthonasal and retronasal odorant stimulation: functional<br />

insights through complementary techniques<br />

Michelle R Rebello 1,2 , Basavaraju G Sanganahalli 3 , Gordon M<br />

Shepherd 1,2 , Fahmeed Hyder 3 , Justus V Verhagen 1,2<br />

1<br />

The John B. Pierce Laboratory New Haven, CT, USA, 2 Yale School of<br />

Medicine, Dept. Neurobiology New Haven, CT, USA, 3 Yale University,<br />

MRRC New Haven, CT, USA<br />

POSTER PRESENTATIONS<br />

Various techniques can be used to evaluate odor response maps<br />

of the olfactory bulb, each with its own advantages. Here we<br />

combined fMRI with intrinsic and calcium optical imaging to<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

54


etter characterize glomerular responses. Functional MRI and<br />

intrinsic imaging share slow responses of complex origin which<br />

are based on dynamic changes in oxy- and deoxy-hemoglobin.<br />

Whereas fMRI can image the entire bulb, intrinsic and calcium<br />

are limited to the dorsal regions but with high spatiotemporal<br />

resolution. CBV-weighted fMRI was unsuccessful in bulb<br />

imaging, as was the use of other anesthetics (Ex:domitor). In<br />

seven rats we per<strong>for</strong>med micro fMRI (BOLD, 120x120x300µm)<br />

in dorsal orientation so that data could be compared to<br />

subsequent calcium imaging of orthonasal responses in the<br />

same subjects. In a few cases odor induced activation maps<br />

showed strong overlap between the two imaging modalities.<br />

We further apply these techniques in separate animals to define<br />

retronasal dorsal and whole bulb responses and compare those<br />

to orthonasal odorant presentations. We have found significant<br />

effects of odor route on the response magnitude and timing.<br />

Further, retronasal odor concentration affects response latency<br />

in a direction depending on the odorant. We are evaluating<br />

several methods of post-hoc co-registering across these methods<br />

1) across functional maps (fMRI, intrinsic, calcium), 2) using<br />

phantom markers on the skull, 3) using vasculature via venogram<br />

MRI, and 4) using SWIFT-MRI <strong>for</strong> skull imaging. Due to the<br />

partial volume effect and the relatively thin dorsal glomerular<br />

layer we are also developing a miniature phased coil-array<br />

allowing higher resolution and high quality dorsal functional<br />

images. Our comparative approach shows the challenges<br />

in obtaining and interpreting odor maps using different<br />

methodologies. Acknowledgements: This work is supported by<br />

NIH/NIDCD Grants R01DC009994 and R01DC011286.<br />

#P68 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Quantifying Bursting Olfactory Neuron Activity from Calcium<br />

Signals Using Maximum Entropy Deconvolution<br />

In Jun Park 1 , Yuriy V. Bobkov 2 , Barry W. Ache 2,3 , Jose C. Principe 1<br />

1<br />

Dept. of Electrical and Computer Engineering, University of Florida<br />

Gainesville, FL, USA, 2 Whitney Laboratory, Center <strong>for</strong> Smell and<br />

Taste, and McKnight Brain Institute St Augustine, FL, USA, 3 Depts. of<br />

Biology and Neuroscience, University of Florida Gainesville, FL, USA<br />

Advances in calcium imaging have enabled studies of the activity<br />

dynamics of both individual neurons and neuronal assemblies.<br />

However, inferring action potentials (spikes) from calcium signals<br />

is still a challenging issue due to hidden nonlinearity in their<br />

relationship, contamination by noise, and often the relatively<br />

low temporal resolution of the calcium signal compared to the<br />

time-scale of spike generation. Complex neuronal discharge,<br />

as in the case of the bursting or rhythmically active neuronal<br />

activity represents an even greater challenge <strong>for</strong> reconstructing<br />

spike trains based on calcium signals. Here we propose doing this<br />

using blind calcium signal deconvolution based on a theoretical<br />

in<strong>for</strong>mation approach. The basic idea is to maximize the output<br />

entropy of a nonlinear filter where the nonlinearity is defined<br />

by the cumulative distribution function of the spike signal.<br />

We tested this maximum entropy (ME) algorithm on bursting<br />

olfactory receptor neurons (bORNs) in the lobster olfactory<br />

organ. The advantage of the ME algorithm is that the filter can<br />

be trained online based only on the statistics of the spike signal<br />

without making any assumptions about the spike-calcium signal<br />

relation. We show that the ME method is able to reconstruct<br />

the timing of the first and the last spike of a burst with higher<br />

accuracy compared to other methods. Thus the ME method<br />

should be a useful tool <strong>for</strong> inferring parameters of bursting<br />

neurons, including bursting olfactory neurons, to help further<br />

understand the mechanism and function of bursting-based<br />

neuronal sensory coding. Acknowledgements: Supported by<br />

award R21 DC011859 from the NIDCD<br />

#P69 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Allometric Growth of Olfactory Bulb and Brain in Female<br />

Minks<br />

Willi Bennegger 1 , Elke Weiler 1,2<br />

1<br />

Maria-von-Linden-Schule, Heckentalstraße 86 89518 Heidenheim,<br />

Germany, 2 Faculty of Medicine, Institute of Anatomy,<br />

Department of Neuroimmunology, University of Leipzig, Liebigstr.<br />

13 04103 Leipzig, Germany<br />

The olfactory bulb is the anterior part of the brain and<br />

phylogenetically one of the oldest brain structures. During<br />

postnatal development, when the animal grows, the brain<br />

increases in size – and so does the olfactory bulb. However, in<br />

some mammals, such as the American mink (Neovison vison) it is<br />

known, that brain shows an overshoot development postnatally<br />

with a subsequent reduction in size. Thus we were interested,<br />

if this applies also to the olfactory bulb. There<strong>for</strong>e we analyzed<br />

morphometrically a total of 57 female minks ranging from<br />

newborn (postnatal day 0, P0) to one year of age <strong>for</strong> their brain<br />

and olfactory bulb size. The results reveal, that the volume of<br />

one olfactory bulb in newborns is 1.26 ± 0.02 mm 3 , increasing<br />

continuously (P30: 40.27 ± 8.77 mm 3 ; P90: 84.70 ± 1.87 mm 3 ) to<br />

adult values (107.19 ± 4.15 mm 3 ) with no overshoot phenomena.<br />

In contrast, the brain weight increases postnatally from P0 (0.29<br />

± 0.06 g) up to P90 (10.18 ± 0.42 g) when maximal values are<br />

reached, and decreasing afterwards more than 17% to the adult<br />

size (8.43 ± 0.35 g). The olfactory bulb growth there<strong>for</strong>e does not<br />

parallel the total brain growth but shows an allometric growth<br />

pattern. On the other hand, the overall body growth increases<br />

continuously to adult values resulting in an olfactory bulb/<br />

body weight ratio of similar values among newborns, brain<br />

overshoot age and adults (P0: 0.014±0.002 %; P90: 0.012±0.002<br />

%, adult: 0.011±0.001 %) with higher values early postnatally<br />

(P30: 0.047±0.013 %). This indicates that the olfactory bulb<br />

is influenced by other factors than the cortical neurons <strong>for</strong> its<br />

neuronal network growth and controlled by different stimuli <strong>for</strong><br />

its <strong>for</strong>mation and connectivity. Further, no postnatal reduction<br />

in size suggests a basic and important functional relevance of the<br />

olfactory bulb.<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

55


#P70 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Effect of kamikihito (TJ-137) on nerve growth factor and<br />

olfactory nerve in vivo<br />

Junpei Yamamoto 1 , HIdeaki Shiga 1 , Kohshin Washiyama 2 ,<br />

Ryohei Amano 2 , Takaki MIwa 1<br />

1<br />

Otorhinolaryngology, Kanazawa Medical University Ishikawa, Japan,<br />

2<br />

Quantum medical technology, Kanazawa University Ishikawa, Japan<br />

Background: A Kampo product, kamikihito (product name code:<br />

TJ-137), has ingredients that promote nerve growth. Paclitaxel,<br />

a cancer chemotherapeutic agent, is toxic to olfactory nerve cells<br />

in vivo. To show TJ-137 pre-medication effect on nerve growth<br />

factor (NGF) in olfactory bulb and paclitaxel induced olfactory<br />

neuropathy in vivo. Methods: Female 7-week-old Bulb/c mice<br />

were fed food containing TJ-137, or control food, <strong>for</strong> 14 days<br />

be<strong>for</strong>e and after intravenous paclitaxel administration. NGF<br />

was assessed with ELISA in the olfactory bulb of mice fed<br />

food containing TJ-137 (n=9), or control food (n=9) <strong>for</strong> 14<br />

days. The epithelial changes in the nasal turbinates of mice<br />

were assessed by H&E and immunohistochemical staining <strong>for</strong><br />

the olfactory marker protein (OMP). The accumulation of the<br />

neuronal tracer (Dextran tetramethylrhodamine) in the olfactory<br />

bulb was assessed in frozen sections of mice 48 h after nasal<br />

administration of the tracer. Results: NGF was significantly<br />

increased in the olfactory bulb of mice fed food-containing<br />

TJ-137 than in the control mice (P=0.017). The epithelium<br />

of nasal turbinates of TJ-137 treated mice was less injured<br />

than that of the control mice after paclitaxel administration.<br />

The accumulation of the neuronal tracer in the olfactory bulb<br />

was higher in the TJ-137 treated mice compared to controls.<br />

Conclusion: We found that TJ-137 is effective in increasing<br />

NGF in olfactory bulb and reducing paclitaxel induced olfactory<br />

neuropathy in vivo. Acknowledgements: Assist Kaken from<br />

Kanazawa Medical University (2012)<br />

#P71 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Fgf8 Defines Neurogenic Vomeronasal and GnRH Neurogenic<br />

Mileu by Influencing BMP and BMP Antagonists Expression.<br />

Paolo E. Forni 1 , Kapil Bharti 2 , Susan Wray 1<br />

1<br />

National Institute of Neurological Disorders and Stroke/NIH<br />

Bethesda, MD, USA, 2 National Eye Institute/NIH Bethesda,<br />

MD, USA<br />

FGF8 plays a pivotal role in development of craniofacial<br />

structures, the olfactory/vomeronasal system and GnRH<br />

neurons. BMPs can antagonize FGFs expression and signal,<br />

BMP and FGF8 signaling are known to exert opposite roles<br />

in defining epithelial versus neurogenic fate. We analyzed<br />

the relation between Fgf8, BMP and BMP anatgonists in the<br />

developing olfactory pit in two Fgf8 hypomorph mouse models,<br />

expressing different levels of FGF8. In both mutant mouse<br />

models, Fgf8 neo/neo and Fgf8 neo/Null , regardless of the FGF8<br />

dosage, overlapping defects were observed in the olfactory pit:<br />

lack of neurons <strong>for</strong>mation limited to the ventral area of the<br />

developing nasal pit and the proximal portion of respiratory<br />

epithelium, where GnRH neurons normally <strong>for</strong>m. Analyzing<br />

expression of BMP4 and BMP antagonist Noggin we found<br />

a previously not described large mesenchymal Noggin source<br />

that sharply defines a BMP free GnRH and VNO neurogenic<br />

border. In Fgf8 hypomorphs BMP4 and Noggin expression were<br />

found to be altered, with Noggin no longer defining the GnRH<br />

and VNO neurogenic border. These data suggest that the role<br />

played by FGF8 in controlling cell fate specification and neural<br />

patterning of the olfactory pit is in large part indirect as Fgf8<br />

levels affect both BMP and Noggin expression in the pit and<br />

nasal mesenchyme. BMP and Noggin expression respectively<br />

define the epithelial or neurogenic permissive borders. The<br />

previously described lack of GnRH neuron specification in<br />

animals with chronically reduced FGF8 reflects the loss of a<br />

large neurogenic permissive milieu that includes the entire VNO.<br />

Acknowledgements: The work presented was supported by<br />

the Intramural Research Program of the National Institutes of<br />

Health, National Institute of Neurological Disorders and Stroke<br />

#P72 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

An IP3R3- and NPY-expressing microvillous cell mediates<br />

tissue homeostasis and regeneration<br />

Colleen C. Hegg 1,2,3 , Cuihong Jia 1 , Sebastien Hayoz 1 ,<br />

Chelsea R. Hutch 2,3 , Apryl E. Pooley 2 , Tania R. Iqbal 2<br />

1<br />

Michigan State University Pharmacology and Toxicology East Lansing,<br />

MI, USA, 2 Neuroscience Program East Lansing, MI, USA,<br />

3<br />

Center <strong>for</strong> Integrative Toxicology East Lansing, MI, USA<br />

Calcium-dependent release of neurotrophic factors plays an<br />

important role in the maintenance of neurons, yet the release<br />

mechanisms are understudied. The inositol triphosphate (IP3)<br />

receptor is a calcium release channel that has a physiological<br />

role in development, sensory perception, neuronal signaling and<br />

secretion. In the olfactory system, the IP3 receptor subtype 3<br />

(IP3R3) is expressed exclusively in a microvillous cell subtype<br />

that is the predominant cell that expresses neurotrophic factor<br />

neuropeptide Y (NPY). We hypothesized that the IP3R3-<br />

expressing microvillous cells secrete sufficient NPY needed <strong>for</strong><br />

both the continual maintenance of the neuronal population<br />

and <strong>for</strong> neuroregeneration following injury. We addressed this<br />

question by assessing the release of neurotrophic factor NPY<br />

and regenerative capabilities in wild type mice, IP3R3 +/- , and<br />

IP3R3 -/- mice. Injury, simulated using extracellular ATP as a<br />

model, induced IP3 receptor-mediated NPY release in wild-type<br />

mice. ATP-evoked NPY release was impaired in IP3R3 -/- mice,<br />

suggesting that IP3R3 contributes to NPY release following<br />

injury. Under normal physiological conditions, both IP3R3 -<br />

/-<br />

mice and explants from these mice had fewer progenitor<br />

cells that proliferate and differentiate into immature neurons.<br />

Although the number of mature neurons and the in vivo rate of<br />

proliferation were not altered, the proliferative response to the<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

56


olfactotoxicant satratoxin G and olfactory bulb ablation injury<br />

was compromised in the olfactory epithelium of IP3R3 -/- mice.<br />

The reductions in both NPY release and number of progenitor<br />

cells in IP3R3 -/- mice point to a role of the IP3R3 in tissue<br />

homeostasis and neuroregeneration. Collectively, these data<br />

suggest that IP3R3 expressing microvillous cells are actively<br />

responsive to injury and promote recovery. Acknowledgements:<br />

Supported by NIH DC006897 (CCH), MSU institutional funds<br />

(CCH), NIEHS T32 ES007255 (CRH), NINDS T32 NS044928<br />

(AEP and TRI) and Swiss Fellowship <strong>for</strong> Advanced Researchers<br />

PA 00P3_131493 (SH).<br />

#P73 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

PACAP increases [Ca 2+ ] i<br />

in neonatal OB via direct and<br />

indirect mechanisms<br />

Mavis A Irwin 1 , Mary T Lucero 2<br />

1<br />

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

2<br />

American University of the Caribbean Cupecoy, Netherlands Antilles<br />

Our lab has been studying the pleiotropic peptide named<br />

Pituitary Adenylate Cyclase Activating Peptide (PACAP) in<br />

the olfactory epithelium 1 of rodents. The physiological effects<br />

of PACAP in the olfactory bulb (OB) are still unknown.<br />

Neonatal OB is enriched with both PACAP and its G-protein<br />

coupled receptor PAC1R. Without PACAP, neonates often die<br />

be<strong>for</strong>e weaning, suggesting that PACAP is required <strong>for</strong> normal<br />

development. Previously, we showed that PACAP led to an<br />

oscillating increase in [Ca 2+ ] i<br />

in OB neurons. To address whether<br />

the PACAP-induced responses are direct or indirect, we used<br />

cocktails of antagonists <strong>for</strong> the GABA receptors (GABAR) and/<br />

or glutamate receptors (GlutR) in the presence and absence of<br />

PACAP. We per<strong>for</strong>med confocal Ca 2+ imaging on live slices<br />

from P2-P5 mice loaded with the Ca 2+ indicator dye Fluo-4<br />

AM. The optimal dose of PACAP was empirically determined<br />

to be 40 nM and was used in all experiments. Combined block<br />

of GABAR and GlutR yielded a 66% decrease in numbers<br />

of PACAP responsive cells. Blocking just GlutR resulted in<br />

a similar reduction, suggesting that glutamate mediates the<br />

majority of the indirect effects. Interestingly, blocking only<br />

the GABAR resulted in block of GABA-induced initial Ca 2+<br />

response on immature cells. However, the majority of these cells<br />

showed the post-PACAP oscillation. Our data suggest 1) about<br />

1/3 of the PACAP-responsive cells have direct PAC1R activity.<br />

2) PACAP promotes glutamate release which in turn activates<br />

2/3 of the PACAP-responsive cells. 3) GlutR may have a role in<br />

the post-PACAP [Ca 2+ ] oscillation. 4) GABA is also released by<br />

PACAP from PAC1R-rich GABAergic cells. In conclusion, we<br />

find that PACAP has both direct and indirect effects on neonatal<br />

OB neurons and may promote glutamate and GABA release in<br />

early development. Acknowledgements: NIH 1F31DC011686-02<br />

Blackman Trust Fund 00253 6000 16917 University of Utah<br />

Graduate Research Fellowship 2011<br />

#P74 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Inactivation of the Interoceptive Insula Suppresses<br />

Chemosensory Cue Reactivity to Ethanol Following<br />

Chronic Ethanol Exposure<br />

Norma Castro, Emily Driver, Jason Dudley, Jessica Godfrey,<br />

Brian Feretic, Carlo Quintanilla, Susan M. Brasser<br />

San Diego State University/Psychology San Diego, CA, USA<br />

Recent findings indicate that the insular cortex is critically<br />

involved in addictive behavior to multiple drugs of abuse by<br />

regulating an organism’s responsiveness to drug-associated<br />

sensory cues. Damage to the insula in addicted smokers results in<br />

a disruption of nicotine addiction, and inactivation of the insula<br />

in rodents disrupts conditioned preference <strong>for</strong> environments<br />

previously paired with amphetamine or nicotine. Imaging<br />

studies have demonstrated activation in the insula during drug<br />

craving and exposure to drug-related cues, including the taste<br />

of alcohol in heavy drinkers. The present study measured<br />

chemosensory responses to ethanol in chronically ethanolexposed<br />

or naive rats under conditions of pharmacological<br />

silencing of the visceral insula to examine the role of this brain<br />

region in mediating responses to ethanol-associated sensory<br />

cues. Rats were initially exposed to either a 20% ethanol<br />

intermittent access paradigm or were given access only to water.<br />

Following implantation of intracranial cannulae, rats from each<br />

exposure condition were tested <strong>for</strong> brief-access lick responses<br />

to ethanol (0-40%) after receiving bilateral insula infusions<br />

of saline or muscimol. Alcohol-experienced rats displayed a<br />

concentration-dependent increase in chemosensory avidity <strong>for</strong><br />

ethanol compared to alcohol-naive rats, evidenced by elevated<br />

lick responses and trial sampling frequency particularly at higher<br />

ethanol concentrations (15-40%). Inactivation of the insula<br />

eliminated this concentration-dependent response in chronically<br />

exposed animals, but did not modify orosensory responses to<br />

ethanol in alcohol-naive rats. These data support insular cortex<br />

involvement in mediating responsiveness to conditioned alcohol<br />

chemosensory cues following chronic association of ethanol’s<br />

taste and post-absorptive effects. Acknowledgements: Support<br />

Contributed By: NIH AA015741<br />

#P75 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Growth patterns of sensory neuron axon terminals in the<br />

developing olfactory bulb<br />

Ivan Manzini 1,2 , Thomas Hassenlklöver 1,2<br />

1<br />

University of Göttingen, Department of Neurophysiology and Cellular<br />

Biophysics Göttingen, Germany, 2 University of Göttingen, DFG Cluster<br />

of Excellence “Nanoscale Microscopy and Molecular Physiology of the<br />

Brain” (CNMPB) Göttingen, Germany<br />

The developing, but also the mature vertebrate olfactory<br />

system is a site of ongoing neurogenesis. Olfactory stem cells<br />

continuously generate new sensory neurons which extend<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

57


axons into the olfactory bulb where they face the challenge to<br />

integrate into an existing neuronal circuitry. Synaptic contacts<br />

to second-order neurons are <strong>for</strong>med in distinct target regions,<br />

so-called glomeruli. In rodents, sensory neurons normally<br />

project only into one specific glomerulus of the olfactory bulb.<br />

We investigated the growth patterns of sensory neuron axons<br />

in the developing olfactory system of the aquatic amphibian<br />

Xenopus laevis. To address the question how connectivity is<br />

reshaped during olfactory system maturation a range of larval<br />

stages and young postmetamorphic animals were included<br />

in the experiments. Fluophore-coupled dextrans or plasmid<br />

DNA, encoding <strong>for</strong> fluorescent proteins, were introduced<br />

into sensory neurons via electroporation. The main sensory<br />

projection fields within the main- and accessory olfactory bulb<br />

were visualized by electroporation of the whole olfactory organ.<br />

During metamorphosis the main olfactory system is completely<br />

reorganized, whereas the sensory neurons of the accessory<br />

olfactory system are maintained. The axonal branching patterns<br />

of sensory neurons, originating from both the vomeronasal and<br />

main olfactory epithelium, were investigated by sparse staining<br />

of sensory neurons. Synaptic connections were clearly visible as<br />

tufted axonal endings. Most sensory neurons showed a branched<br />

axonal pattern be<strong>for</strong>e terminating in tufted arborizations inside<br />

glomeruli. Surprisingly, a high percentage of cells terminated in<br />

multiple and not single glomerulus-like structures. This pattern<br />

was comparable in sensory neurons originating from both the<br />

vomeronasal and the main olfactory organ. Acknowledgements:<br />

Supported by DFG Cluster of Excellence “Nanoscale<br />

Microscopy and Molecular Physiology of the Brain” (CNMPB)<br />

to I.M. and DFG Schwerpunktprogramm 1392 to I.M.<br />

#P76 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Activity-Dependent Expression of Odorant Receptors<br />

in the Mouse Olfactory Epithelium<br />

Shaohua Zhao 1,2 , Huikai Tian 1 , Rosemary Lewis 1 , Limei Ma 3 ,<br />

Ying Yuan 1 , Congrong R Yu 3 , Minghong Ma 1<br />

1<br />

Department of Neuroscience, University of Pennsylvania School of<br />

Medicine Philadelphia, PA, USA, 2 Department of Geriatric Cardiology,<br />

Qilu Hospital of Shandong University Jinan, China, 3 Stowers Institute<br />

<strong>for</strong> Medical Research Kansas City, MT, USA<br />

Sensory experience plays critical roles in development and<br />

maintenance of the olfactory system, which undergoes<br />

considerable neurogenesis throughout life. In the mouse olfactory<br />

epithelium, each primary olfactory sensory neuron (OSN) stably<br />

expresses a single odorant receptor (OR) type out of a repertoire<br />

of ~1200. All OSNs with the same OR identity are distributed<br />

within one of the few broadly-defined zones. However, it remains<br />

elusive whether such OR expression patterns are shaped by<br />

sensory stimulation and/or neuronal activity. Here we addressed<br />

this question by investigating OR gene or protein expression in<br />

two surgically- or genetically-modified mouse models. Using in<br />

situ hybridization, we examined the expression patterns of 15<br />

selected OR genes in mice which underwent neonatal, unilateral<br />

naris closure. After four-week occlusion, the expression level in<br />

the closed side was significantly lower (<strong>for</strong> four ORs), similar<br />

(<strong>for</strong> three ORs) or significantly higher (<strong>for</strong> eight ORs) than that<br />

in the open side. In addition, using a specific OR antibody,<br />

we demonstrated that this OR protein was upregulated in the<br />

closed side but downregulated in the open side. Furthermore,<br />

we examined the expression patterns of individual OR genes<br />

in transgenic mice in which olfactory marker protein (OMP)<br />

drives overexpression of the inward rectifying potassium channel<br />

(Kir2.1) in most mature OSNs to reduce their neuronal activity.<br />

The cell density <strong>for</strong> most OR genes (six out of seven tested) was<br />

significantly reduced compared to wild-type controls. The results<br />

suggest that sensory inputs have differential influence on OSNs<br />

expressing different ORs and that neuronal activity is critical <strong>for</strong><br />

survival of OSNs. Acknowledgements: Supported by grants from<br />

the NIDCD/NIH DC006213 and DC011554.<br />

#P77 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Optogenetic Investigation of GABAergic Circuitries in the<br />

Rostral Nucleus of the Solitary Tract<br />

James A. Corson, Robert M. Bradley<br />

University of Michigan/ Biologic and Materials <strong>Sciences</strong> Ann Arbor,<br />

MI, USA<br />

The rostral nucleus of the solitary tract (rNTS) is the first central<br />

target of primary gustatory nerve fibers and as such plays an<br />

essential role in the processing and coding of peripheral taste<br />

sensory in<strong>for</strong>mation. The intrinsic circuitry within rNTS is likely<br />

integral in shaping the incoming in<strong>for</strong>mation into both ascending<br />

and descending efferent signals. Substantial subpopulations of<br />

interneurons in the rNTS are GABAergic and thus contribute<br />

to the generation of hyperpolarization-activated changes in<br />

repetitive firing patterns in projection neurons. Despite this<br />

importance in shaping rNTS gustatory-evoked signaling, the<br />

organization of rNTS GABAergic circuits is unknown. To<br />

investigate the organization of GABAergic innervation onto<br />

identified populations of neurons, we used a mouse model<br />

in which channelrhodopsin was expressed under the control<br />

of the vesicular GABA transporter. GABAergic interneurons<br />

were activated in an in vitro slice preparation with 473 nm laser<br />

illumination merged into the optic train of the microscope.<br />

Focused laser illumination produced consistent saturated<br />

photocurrents in GABAergic neurons with high temporal and<br />

spatial resolution. While recording inhibitory postsynaptic<br />

currents in either GABAergic or non-GABAergic neurons, the<br />

laser spot was systematically scanned over discrete portions<br />

of the rNTS to map out the GABAergic innervation onto the<br />

recorded neuron. Neurons received inhibitory innervation<br />

from wide expanses of rNTS, often with focal spots of strong<br />

inhibition located in areas not immediately adjacent to the<br />

recorded neuron. This suggests that in addition to a low level<br />

of global inhibition, there are also specific subregions of rNTS<br />

that are able to strongly hyperpolarize individual neurons and<br />

possibly induce alterations in repetitive discharge patterns.<br />

Acknowledgements: T32DC000011, RO1DC000288<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

58


#P78 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Sensory Afferents from the Stomach of the Rat Converge<br />

onto Taste-responsive Neurons in the Rat Brainstem<br />

Alexander J. Denman-Brice, Patricia M. Di Lorenzo<br />

Binghamton University/Psychology Binghamton, NY, USA<br />

Recent descriptions of taste receptors in the gut have set the stage<br />

<strong>for</strong> the idea that taste stimuli continue to provide in<strong>for</strong>mation<br />

even when they are no longer in the mouth. For example,<br />

intraduodenal infusions of bitter tastants can modify eating<br />

behavior within a single meal. Given that both taste and postlingual<br />

chemosensory feedback may be important <strong>for</strong> modifying<br />

eating behavior on a relatively short timescale, it is possible<br />

that chemosensory afferents from the gut may converge onto<br />

the same relay nuclei as taste in<strong>for</strong>mation in the brainstem.<br />

We investigated the responsiveness of single neurons in the rat<br />

brainstem to tongue and gastric taste stimulation. Initially, rats<br />

were anesthetized with urethane and prepared <strong>for</strong> recording from<br />

the brainstem. A length of polyethylene tubing was threaded<br />

down the rat’s esophagus to the stomach <strong>for</strong> delivery of tastants.<br />

A tungsten microelectrode was then placed in the nucleus of<br />

the solitary tract (NTS) and taste-responsive cells were isolated.<br />

Preliminary data from 18 taste-responsive cells show that some<br />

(n=8) NTS cells change their firing rate in response to infusion<br />

of small amounts (0.4 ml) of taste stimuli (0.1 M NaCl, 0.01 M<br />

HCl, 0.01 M quinine, 0.5 M sucrose and 0.1 M MSG) into the<br />

stomach. Gastric responses were most frequently found to NaCl<br />

and MSG; no excitatory responses were found to HCl infused<br />

into the stomach. An additional cell was not responsive to<br />

lingual taste stimuli but was inhibited by gastric tastant delivery<br />

of MSG, HCl and quinine. Collectively, these data suggest that<br />

in<strong>for</strong>mation from both lingual and post-lingual chemoreceptors<br />

converge onto NTS cells, suggesting a role <strong>for</strong> post-lingual<br />

chemoreceptors in modulation of ingestive behavior on a short<br />

timescale. Acknowledgements: Supported by NIDCD grant<br />

RO1DC006914 to PMD.<br />

#P79 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Synaptic Properties of the Basolateral Amygdala Projection<br />

to Gustatory Cortex<br />

Melissa Haley 1,2 , Alfredo Fontanini 1,2 , Arianna Maffei 1,2<br />

1<br />

Department of Neurobiology and Behavior Stony Brook, NY, USA,<br />

2<br />

Program in Neuroscience Stony Brook, NY, USA<br />

The gustatory cortex (GC) receives a dense and anatomically<br />

well documented projection from the basolateral nucleus of the<br />

amygdala (BLA). This input is thought to convey affective and<br />

anticipatory in<strong>for</strong>mation regarding taste. Indeed inactivation<br />

of BLA results in dramatic changes in the way taste and<br />

anticipatory cues are processed in GC. Although the behavioral<br />

and functional significance of this projection has been the focus<br />

of extensive investigation in awake and anesthetized animals,<br />

the synaptic properties and organization of these inputs are<br />

unknown. To study the BLA to GC synapse, viral vectors<br />

carrying a construct <strong>for</strong> ChannelRhodopsin2 were injected in<br />

the BLA. After 2 weeks, whole-cell recordings in dysgranular<br />

and agranular GC were per<strong>for</strong>med in combination with<br />

photoactivation of BLA terminal fields. BLA afferents were<br />

found to target both excitatory and inhibitory neurons in all<br />

layers of GC. Across all layers, approximately 60% of regularspiking<br />

(RS), fast-spiking (FS), and low-threshold spiking (LTS)<br />

neurons responded to light-activation of BLA afferents. RS cells<br />

had a longer rise time than FS cells (p=.016) and a longer decay<br />

time than FS cells (p=.0004). In addition, differences could be<br />

seen in the synaptic properties of BLA inputs onto neurons in<br />

superficial and deep layers of GC. Layer 2/3 RS cells had larger<br />

current amplitudes than layer 5/6 cells (p=.03), and there was a<br />

significant difference in the percentage of FS cells that responded<br />

to stimulation in layer 2/3 (75%) and layer 5/6 (25%). These data<br />

suggest that BLA inputs in GC have cell type specific and layer<br />

specific properties. The combination of feed-<strong>for</strong>ward inhibition<br />

and excitation likely serves to shape the temporal dynamics of<br />

taste responses and to enhance the representation of behaviorally<br />

salient stimuli. Acknowledgements: National Eye Institute Grant<br />

#R01-EY019885-S1 and National Institute on Deafness and<br />

Other Communication Disorders Grant #R01-DC010389<br />

#P80 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Cortical modulation of taste-related orofacial behaviors<br />

Jennifer X. Li, Donald B. Katz<br />

Brandeis University Waltham, MA, USA<br />

Upon receiving a taste stimulus, primates and rodents produce a<br />

sequence of rhythmic orofacial movements, also known as taste<br />

reactivity (TR), specific elements of which reflect the hedonic<br />

quality (i. e., the palatability) of the stimulus. Aversive stimuli,<br />

such as quinine, elicit bouts of gapes, movements that serve to<br />

eject the stimulus from the mouth. The production of gapes<br />

indicates aversiveness, and the latency to gape is inversely related<br />

to the degree of aversiveness (Travers and Norgren, 1986). While<br />

the motor circuits necessary <strong>for</strong> taste-related orofacial movements<br />

are contained within a brainstem network (Grill & Norgren), in<br />

the intact animal, this network is modulated by feedback from<br />

higher-order <strong>for</strong>ebrain structures. Relatively little is known about<br />

<strong>for</strong>ebrain structures’ roles in the selection and production of<br />

TR, however. We set out to examine the relationship between<br />

TR and neural responses in primary gustatory cortex (GC) by<br />

per<strong>for</strong>ming paired recordings of single unit activity and jaw<br />

electromyography (EMG) in awake rats presented with strong<br />

(0.3 M) and weak (0.03 M) sucrose, and strong (0.001 M) and<br />

weak (0.0001 M) quinine, via intra-oral cannulae. Comparisons<br />

of the time courses of neural and EMG responses revealed that<br />

palatability-related signals in GC preceded those in EMG by<br />

~250 ms, suggesting that GC could drive palatability-related<br />

orofacial movements. Preliminary analyses further demonstrated<br />

that the spiking activity of individual GC neurons was correlated<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

59


with the latency of quinine-induced gape bouts. Our results<br />

suggest that <strong>for</strong>ebrain activity may influence certain aspects<br />

of TR, such as the initiation of palatability-related orofacial<br />

movements. Acknowledgements: DC007703<br />

#P81 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Neural dynamics in response to binary taste mixtures<br />

Joost X Maier, Donald B Katz<br />

Brandeis University Waltham, MA, USA<br />

In natural environments, taste signals an animal encounters<br />

typically consist of complex mixtures of tastants. Although a<br />

great deal is known about how the taste system processes single<br />

tastes presented in isolation, not much is know about how brain<br />

integrates different taste signals presented simultaneously. Here<br />

we probed single neurons in primary gustatory cortex (GC)<br />

<strong>for</strong> responsiveness to binary taste mixtures. Stimuli consisted<br />

of sucrose and citric acid and sucrose and sodium chloride<br />

mixed in different ratios (100%/0%, 90%/10%, 70%/30%,<br />

50%/50%, etc.). We tested <strong>for</strong> three different hypothetical<br />

response patterns: 1) Responses varying as a function of sucrose<br />

concentration (the monotonic pattern); 2) Responses increasing<br />

or decreasing as a function of degree of mixture of the stimulus<br />

(the mixture pattern); and 3) Responses that change abruptly<br />

from being similar to one pure taste to being similar the other<br />

(the categorical pattern). Our results demonstrate the presence<br />

of both monotonic and mixture patterns within responses<br />

of GC neurons. Specifically, further analysis (that included<br />

the presentation of 50 mM sucrose and citric acid) made it<br />

clear that mixture suppression reliably precedes a palatabilityrelated<br />

pattern, and that the same phenomenon characterizes<br />

responses to sucrose/NaCl mixtures. The temporal dynamics<br />

of the emergence of the palatability-related pattern parallel the<br />

temporal dynamics of the emergence of preference behavior <strong>for</strong><br />

the same mixtures, as measured by a brief access test. We saw no<br />

evidence of categorical coding.<br />

#P82 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Conditioned Taste Aversion Does Not Require Cortical<br />

mRNA Synthesis<br />

Abigail A Russo 1 , Yasmin U Marrero 2 , Donald B Katz 1,2,3<br />

1<br />

Brandeis University Department of Psychology Waltham, MA, USA,<br />

2<br />

Brandeis University Program of Neuroscience Waltham, MA, USA,<br />

3<br />

Volen Center <strong>for</strong> Complex Systems, Brandeis University Waltham,<br />

MA, USA<br />

Although it has been well established that the gustatory cortex<br />

(GC) plays a significant role in the consolidation of taste<br />

memory, the precise physiological mechanisms by which this<br />

takes place are not fully understood. Notably, taste memory<br />

acquisition is traditionally viewed as dependent on cortical<br />

protein synthesis (Dudai et al. 2004), but it is unclear whether<br />

the learning process requires cortical mRNA transcription.<br />

Here, we investigated this possibility using actinomycin D<br />

(Act-D), an mRNA synthesis inhibitor that has been shown<br />

to impair contextual fear conditioning when infused into the<br />

amygdala (Parsons et al 2006). Act-D was microinfused via<br />

cannulae implanted into the GC (1.4 mm anterior to Bregma,<br />

5.0 mm lateral, 4.5 mm ventral) of awake rats. Immediately after<br />

infusion, a conditioned taste aversion protocol was per<strong>for</strong>med<br />

during which the rats were exposed to a taste paired with<br />

malaise. Our results indicated that rats <strong>for</strong>m taste aversions even<br />

when mRNA synthesis in GC is blocked. These results suggest<br />

that memory consolidation is in part independent of mRNA<br />

synthesis in the gustatory cortex. It is likely that subcortical<br />

production of mRNA, presumably in the amygdala, is sufficient<br />

to support cortical protein synthesis and establish taste memory.<br />

Acknowledgements: R01 DC-006666/DC/NIDCD NIH HHS/<br />

United States<br />

#P83 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Nutritive value, not taste, is necessary <strong>for</strong> flavor<br />

preferences in mice<br />

Jennifer M Strat<strong>for</strong>d<br />

Rocky Mountain Taste & Smell Center, Department of Cell and<br />

Developmental Biology, Neuroscience Program, University<br />

Colorado School of Medicine Aurora, CO, USA<br />

The preference <strong>for</strong> food is dependent primarily upon the<br />

interplay between oral and post-oral factors. The relative<br />

contribution of these two systems to food intake is not fully<br />

explored. Mice that lack the ability to taste, by genetic deletion<br />

of the P2X2/P2X3 purinergic receptor subunits (P2X-KO),<br />

can <strong>for</strong>m a preference <strong>for</strong> monosodium glutamate (MSG) using<br />

only post-ingestive cues. The neural mechanisms that underlie<br />

this ability remain unknown but likely involve viscerosensory<br />

detection of the nutritive qualities of MSG. Thus, the current<br />

study assessed if P2X-KO mice can <strong>for</strong>m a preference <strong>for</strong> the<br />

calorie-free sweetener, SC45647 (SC), which, like MSG, is<br />

appetitive to WT animals. WT and P2X-KO mice were given<br />

training sessions with a flavor alone (e.g. cherry) or with a<br />

different flavor (e.g. grape) mixed with 0.05 mM SC. Then all<br />

animals were given 2-bottle preference tests with both flavors<br />

without SC. During training, WT animals drank more SC<br />

than did P2X-KO mice, suggesting that WT, but not P2X-KO<br />

mice, can taste SC. However, neither WT nor P2X-KO animals<br />

preferred the flavor that was previously paired with SC in flavor<br />

alone preference tests. SC-evoked brain activation was measured<br />

by expression of the immediate early gene c-Fos (cFLI) in the<br />

nuc. solitary tract (nTS)- the primary taste/viscerosensory<br />

nucleus. As previously reported <strong>for</strong> MSG stimulation, SCinduced<br />

cFLI in gustatory (rostral) nTS was less in P2X-KO<br />

animals compared to WT controls. In viscerosensory (caudal)<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

60


nTS, SC-induced cFLI did not differ between WT and P2X-<br />

KO mice. Further, within caudal nTS, SC was less effective<br />

than MSG in evoking cFLI in both lines. Together, these results<br />

suggest that nutritive content, not taste, is necessary to drive food<br />

preferences and that this in<strong>for</strong>mation is represented in the caudal<br />

nTS. Acknowledgements: Supported by NIH grants to JMS and<br />

Thomas E. Finger (U. of Colorado Denver Medical School,<br />

Rocky Mountain Taste and Smell Center, Aurora, CO).<br />

#P84 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Distinct groups of cilia in rat rostral nucleus of the solitary<br />

tract (rNST) labeled with ACIII and Arl13b<br />

Min Wang, Charlotte C. Mistretta, Robert M. Bradley<br />

Biologic & Materials <strong>Sciences</strong>, School of Dentistry University<br />

of Michigan, MI, USA<br />

All chemosensitive in<strong>for</strong>mation derived from stimulation of taste<br />

receptors in the oro-pharynx relays via the rNST in the brain<br />

stem. Neurons of the rNST have been defined using anatomical<br />

and histochemical criteria but recently we have also shown that<br />

many rNST neurons possess primary cilia, small organelles<br />

extending from the cell surface. Primary cilia have been shown to<br />

play an important role during development and are involved in<br />

cell signaling pathways. The importance of cilia in development<br />

is evident in various developmental brain disorders, or<br />

ciliopathies, resulting from disrupted cilia function. We studied<br />

the location of primary cilia in rNST cells in postnatal rat<br />

because they may play a role in signaling during taste processing.<br />

A number of markers have been used to identify primary cilia.<br />

Two widely used markers are ACIII, part of a cAMP pathway,<br />

and Arl13b, part of a cGMP signaling pathway. Previously we<br />

reported that ACIII-labeled primary cilia are present in about<br />

half of rNST neurons. Somatostatin receptor 3 (SSTR3) and<br />

melanin-concentrating hormone receptor 1(Mch1R) co-localize<br />

with the ACIII-labeled cilia. Interestingly, though, Arl13b and<br />

ACIII-labeled cilia are found on different rNST cells. Neither<br />

SSTR3 nor Mch1R co-localizes with Arl13b. In addition,<br />

recently we detected a primary cilium in almost every rNST<br />

astrocyte using an Arl13b antibody, but not ACIII, further<br />

demonstrating differences in rNST primary cilia. Since ACIII<br />

couples with a cAMP pathway and Arl13b couples with a cGMP<br />

signaling pathway it is possible that the cell types that express<br />

different primary cilia play different roles in rNST function.<br />

Acknowledgements: NIH NIDCD Grant DC000288<br />

#P85 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Taste responses change over consecutive days in single<br />

cells in the rat brainstem recorded in the awake animal<br />

Michael S. Weiss 1 , Andrew M. Fooden 1 , Jonathan D. Victor 2 ,<br />

Patricia M. Di Lorenzo 1<br />

1<br />

Binghamton Univeristy Dept. of Psychology Binghamton, NY, USA,<br />

2<br />

Weill Cornell Medical College Dept. of Neurology and Neuroscience<br />

New York, NY, USA<br />

Theories of taste coding have relied on recordings from single<br />

cells in a single session; i.e. snapshots of activity. In contrast,<br />

there is evidence that taste responses can change over days (e.g.<br />

in chorda tympani fibers, Shimatani et al., Physiol. & Behav.,<br />

80(2-3), 309-15, 2003). Here, we present data showing that<br />

taste responses in individual cells in the nucleus of the solitary<br />

tract (NTS) and parabrachial nucleus of the pons (PbN) vary<br />

significantly across consecutive days. Rats were surgically<br />

implanted with a chronic microwire assembly into the NTS<br />

or PbN, allowed to recover, and water deprived. Rats had free<br />

access to a lick spout that delivered taste stimuli or water while<br />

cellular activity was recorded. Thus far, in the NTS, 8 animals<br />

yielded multi-day recordings (range = 2-5 d; median = 2 d); in<br />

the PbN, 5 animals yielded multi-day recordings (range = 2-7<br />

d; median = 2.5 d). To determine whether the recordings on<br />

successive days were likely to represent recordings of the same<br />

neuron, we examined the similarity of the recorded wave<strong>for</strong>m<br />

templates. For 76% of multi-day NTS recordings and 30%<br />

of multi-day PbN recordings, wave<strong>for</strong>ms were highly similar<br />

(wave<strong>for</strong>m template correlation > 0.99). As a control, this degree<br />

of similarity was rare (1.3% of pairs in NTS,


allows visualization and quantification of the interneuron cell<br />

bodies throughout the bulbar layers of the AOB (Krosnowski<br />

et al, 2012). We found significant differences in the number of<br />

cholinergic interneurons in the anterior and posterior glomerular<br />

layer (GL) and the external plexi<strong>for</strong>m layer (EPL), with the<br />

highest numbers of cholinergic interneurons in the anterior GL<br />

and posterior EPL. In the posterior EPL, we also noted a high<br />

density of GFP+ cells <strong>for</strong>ming a ring around the outer edges<br />

of the layer, thus creating a heavy GFP+ population along<br />

the border between the anterior and posterior AOB. We then<br />

examined the possible role of ChAT interneurons in the AOB<br />

using adult male mice exposed to either bedding from a mated<br />

pair or to a male aggressor mouse. We monitored the activity<br />

marker, c-fos, in these regions and found significantly higher<br />

levels of activation of ChAT-expressing cells in the EPL in both<br />

exposure groups when compared to activation in control mice,<br />

suggesting that this population may serve a role in the processing<br />

of social odor cues. Thus, we have identified a significant<br />

cholinergic interneuron population in the AOB that varies<br />

significantly in the anterior and posterior regions. Thus, our data<br />

supports the idea that the anterior and posterior AOB process<br />

sensory in<strong>for</strong>mation differently and suggests that this cholinergic<br />

interneuron population may serve to process olfactory<br />

in<strong>for</strong>mation in a region specific manner. Acknowledgements:<br />

Supported by research grants NIH/NIDCD 009269, 012831 and<br />

ARRA administrative supplement to WL<br />

#P87 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Suppression of <strong>Association</strong> Synapses in Piri<strong>for</strong>m Cortex<br />

During Post-Training Sleep Impairs Odor Memory Selectivity<br />

Dylan C Barnes 1,2 , Donald A Wilson 1,2,3<br />

1<br />

Graduate Center CUNY New York City, NY, USA, 2 Nathan Kline<br />

Institute Orangeburg, NY, USA, 3 NYU Langone Medical Center New<br />

York City, NY, USA<br />

Slow wave sleep (SWS) is characterized by slow-wave<br />

oscillations in neocortex, as well as sharp waves (SPW) in<br />

both the hippocampus and piri<strong>for</strong>m cortex (PCX). Neural<br />

activity during SWS is hypothesized to contribute to memory<br />

consolidation through “replay” of waking activity patterns. For<br />

example, we have demonstrated that imposed replay of odorevoked<br />

activity in the olfactory system during SWS enhances<br />

subsequent memory of that odor. Neurons co-activated by an<br />

odor are hypothesized to become linked into a cohesive ensemble<br />

through strengthening of association synapses. Replay of odor<br />

evoked ensemble activity during SWS may help strengthen<br />

these connections and improve memory and memory acuity.<br />

Here, we tested the hypothesis of that association fiber activity<br />

during SWS facilitates replay and memory of recently learned<br />

odors by infusing baclofen (or saline) into the PCX during posttraining<br />

sleep. Baclofen is a GABA-B receptor agonist that has<br />

been shown to selectively depress association fiber synapses.<br />

Rats were chronically implanted with bilateral cannulae and<br />

a recording electrode in the anterior PCX. After recovery, rats<br />

were differentially conditioned with CS+ odor/footshock and<br />

CS- odor stimuli. During the 4 hours immediately post-training,<br />

animals were placed in a sleeping chamber and bilaterally<br />

infused with either baclofen or saline. Local field potential<br />

and EMG activity were recorded during conditioning,<br />

post-training sleep, and test periods. On test day, 24 hours<br />

following conditioning, freezing responses to the CS+, CSand<br />

other odors were examined. Preliminary behavioral results<br />

suggest that post-training PCX baclofen infusions do not<br />

impair memory <strong>for</strong> the CS+ but reduce odor acuity/enhance<br />

generalization of the odor-fear response. Acknowledgements:<br />

F31-DC012284 to D.C.B. and R01-DC003906 from the<br />

NIDCD to D.A.W.<br />

#P88 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Cholinergic modulation of glomerular odor sensitivity<br />

in the olfactory bulb<br />

Mounir Bendahmane, M Cameron Ogg, Max L Fletcher<br />

University of Tennessee Health Science Center Memphis, TN, USA<br />

In the olfactory system, many studies have shown that<br />

cholinergic input to the olfactory bulb is not only involved in<br />

learning and memory but also detection and discrimination.<br />

In this study we used calcium imaging to explore the cholinergic<br />

effect on OB postsynaptic glomerular odor responses. Using mice<br />

expressing GCaMP2 in M/T cells, we studied the modulation<br />

of dorsal surface glomerular odor concentration-response<br />

curves via HDB (horizontal limb of the diagonal band of Broca)<br />

stimulation or OB cholinergic pharmacological manipulation.<br />

Overall, we find that increased cholinergic OB activation through<br />

HDB stimulation or cholinergic-uptake blocker application<br />

increases the sensitivity of individual glomerular odor responses<br />

by shifting the odor concentration-response curve to the left and<br />

decreasing the EC50 by up to one log unit in odor concentration.<br />

This effect was observed <strong>for</strong> all glomeruli tested regardless<br />

of baseline odor sensitivity or odorant used. OB application<br />

of a muscarinic antagonist completely blocks these shifts,<br />

suggesting that the increased sensitivity observed is primarily<br />

driven by muscarinic activation. We are now exploring the<br />

cholinergic effects on individual OB cell types using two-photon<br />

microscopy to further address these effects at the single cell level.<br />

Acknowledgements: NIH R03 DC009853 and the Pew Scholars<br />

Program in the Biomedical <strong>Sciences</strong>.<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

62


#P89 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

#P90 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

In Vivo Optophysiological Analysis of the Glomerular<br />

Unit Response in Mice<br />

Oliver R. Braubach 1,2,3 , Tuce Tombaz 1 , Masoud Allahverdizadeh 1 ,<br />

Thomas Bozza 4,5 , Lawrence B. Cohen 1,2,3 , Ryota Homma 2,3,6<br />

1<br />

Korea Institute of Science and Technology/Center <strong>for</strong> Functional<br />

Connectomics Seoul, Korea, 2 Marine Biological Laboratory Woods<br />

Hole, MA, USA, 3 Yale School of Medicine/Department of Physiology<br />

New Haven, CT, USA, 4 Northwestern University/Department of<br />

Neurobiology Evanston, IL, USA, 5 HHMI Janelia Farm Research<br />

Campus/Visiting Scientist Program Ashburn, VA, USA, 6 The<br />

University of Texas Medical School/Department of Neurobiology and<br />

Anatomy Houston, TX, USA<br />

Olfactory bulb glomeruli organize and relay sensory in<strong>for</strong>mation<br />

that arrives from the nose. Odors typically evoke activity across<br />

many glomeruli, making it difficult to study their individual role<br />

in olfactory processing. To overcome this difficulty we employed<br />

transgenic mice in which channelrhodopsin-2 is selectively<br />

expressed in genetically identified olfactory sensory neurons;<br />

these neurons project their axons to a defined glomerulus<br />

and can be activated via pulsed illumination of the olfactory<br />

epithelium with a 447nm laser. Through in vivo two-photon<br />

calcium imaging, we monitored optically evoked neural activity<br />

in the glomerulus and nearby juxtaglomerular neurons. Laser<br />

pulses reliably activated the glomerulus and evoked calcium<br />

responses in juxtaglomerular neuron pools of various sizes.<br />

Changing laser pulse width and timing altered the strength,<br />

appearance and duration of glomerular and cellular responses.<br />

We also identified juxtaglomerular cells that responded to<br />

optical activation with decreased intracellular calcium; unlike<br />

activated cells, these inhibited cells did not cluster next to the<br />

optically activated glomerulus but were spread throughout<br />

surrounding areas. We are now investigating how different<br />

neuronal responses types relate to the molecular phenotype<br />

of the juxtaglomerular cells. To accomplish this we generated<br />

hybrid mice, which express GAD67-GFP and/or GAD65-<br />

GFP in olfactory bulb interneurons along with the selectively<br />

expressed channelrhodopsin-2. In combination, our data provide<br />

a uniquely detailed analysis of the neuronal network that<br />

comprises a single olfactory glomerulus. Acknowledgements:<br />

Supported by NIH grants U24NS057631 and R01DC005259<br />

and by the National Research Foundation of Korea World Class<br />

Institute Grant WCI 2009-003.<br />

Top-down modulation of olfactory bulb output by the<br />

midbrain serotonergic system.<br />

Daniela Brunert, Matt Wachowiak<br />

Brain Institute and Department of Physiology Salt Lake City, UT, USA<br />

The olfactory bulb (OB) receives input from multiple top-down<br />

neuromodulatory systems. Serotonergic inputs from the raphe<br />

innervate all OB layers and can presynaptically modulate sensory<br />

input gain. However, the effects of serotonergic modulation on<br />

OB circuitry and output in vivo remain unclear. Here, we used<br />

Cre-expressing mouse strains to express the calcium indicator<br />

GCaMP in periglomerular (PG) cells (using GAD2-cre mice)<br />

and in mitral/tufted cells (MTCs) (Cdhr1-cre mice) in order<br />

to visualize how raphe stimulation alters resting and sensoryevoked<br />

excitation of these two neuron populations. In GAD2-<br />

cre mice, brief (1-4 s) electrical stimulation of raphe elicited a<br />

slow increase in baseline fluorescence that outlasted stimulation<br />

by several seconds, as well as a several-fold increase in the<br />

amplitude of inhalation-evoked transients (mean increase, 650<br />

± 247%). Stimulation of raphe also increased odorant-evoked<br />

response amplitudes. Raphe stimulation effects were blocked by<br />

the 5-HT2A/C antagonist cinanserin applied locally to the OB.<br />

These results suggest that serotonergic inputs to OB transiently<br />

increase the baseline excitability of PG cells as well as their<br />

responses to sensory input. In contrast, MTCs in Cdhr1-cre<br />

mice showed only weak increases in baseline fluorescence<br />

upon raphe stimulation onset followed by a more pronounced<br />

increase after stimulation offset in some animals. Surprisingly,<br />

raphe stimulation did not alter inhalation- or odorant-evoked<br />

responses in these neurons. Together, these results demonstrate<br />

a differential effect of serotonergic modulation on OB cell types<br />

in vivo and serve as a starting point <strong>for</strong> further dissection of<br />

the circuit mechanisms underlying the top-down modulation<br />

of early olfactory processing as a function of behavioral state.<br />

Acknowledgements: Funded by NIDCD DC010915<br />

#P91 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Neuronal connections from piri<strong>for</strong>m cortex to prefrontal<br />

cortical areas of mice<br />

Chien-Fu F Chen, Natasha Kharas, Stuart Firestein<br />

Columbia University/Biological <strong>Sciences</strong> New York, NY, USA<br />

Piri<strong>for</strong>m cortex, the main olfactory processing area, has been<br />

shown to have projections to prefrontal cortex providing<br />

olfactory input and receive projections from prefrontal cortex<br />

as a potential downstream modulatory pathway. While recent<br />

data establish the projections from the prefrontal areas to<br />

piri<strong>for</strong>m, the projections from piri<strong>for</strong>m to prefrontal areas<br />

remain less understood. To investigate these connections,<br />

we utilized retrograde tracing and confocal microscopy. We<br />

injected the retrograde tracer, cholera toxin subunit b (CTb),<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

63


via iontophoretic injection in the prefrontal areas of lateral<br />

orbitofrontal cortex (LO) or agranular insular cortex (AI).<br />

The C57BL/6 mice were perfused seven days after CTb injection.<br />

The CTb iontophoretic injection created very confined injection<br />

sites of a diameter of 150 to 200 µm. Our preliminary data<br />

reveal that <strong>for</strong> mice injected in the LO, CTb positive cells were<br />

found in endopiri<strong>for</strong>m nucleus and the ventral-medial area of<br />

layer II/III of the anterior piri<strong>for</strong>m cortex. Furthermore, in the<br />

LO injected mouse, we found labeled cells in agranular insular<br />

and mediodorsal nucleus of thalamus. In the AI injected mice,<br />

a similar labeling pattern was seen in endopiri<strong>for</strong>m nucleus with<br />

sparsely labeled cells in ventral medial anterior piri<strong>for</strong>m cortex.<br />

In both experiments, we found no labeled cells in the dorsal<br />

anterior piri<strong>for</strong>m cortex or posterior piri<strong>for</strong>m cortex. Together<br />

the data suggest that piri<strong>for</strong>m cortex may be composed of finer<br />

anatomical subdivisions that project to prefrontal areas.<br />

Whether these subdivisions per<strong>for</strong>m specific functional roles<br />

remains to be determined.<br />

#P92 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Neuromodulatory regulation of learning within olfactory bulb<br />

Thomas A Cleland<br />

Dept. Psychology, Cornell University Ithaca, NY, USA<br />

Intrinsic plasticity within olfactory bulb (OB) circuitry<br />

modifies odor generalization gradients based on experience<br />

so as to dynamically construct essentially categorical odor<br />

representations. This plasticity also is regulated and perhaps<br />

gated by neuromodulatory state, and can be manipulated by<br />

pharmacological infusions into OB. Cholinergic inputs to the OB<br />

acutely regulate the breadth of odor generalization, though the<br />

nicotinic component, primarily localized in the glomerular layer,<br />

dominates this acute effect. In contrast, muscarinic receptors are<br />

predominantly expressed in the external plexi<strong>for</strong>m layer (EPL),<br />

which has been increasingly associated with mechanisms of<br />

intrinsic learning within OB. We there<strong>for</strong>e investigated the role<br />

that muscarinic modulation of OB circuitry plays in olfactory<br />

associative learning. We found that intrabulbar infusion of<br />

scopolamine impaired olfactory learning when delivered between<br />

training and testing, or when delivered prior to training in studies<br />

imposing a similar 45-minute training-testing latency, but not<br />

when testing followed training immediately or with a fourminute<br />

latency. (Dihydrokainate infusion into OB was used to<br />

prevent the retrograde amnestic effect of isoflurane anesthesia).<br />

This pattern of results indicates that intact muscarinic<br />

responsivity within OB is important <strong>for</strong> the maintenance of an<br />

intact odor memory over this delay period. In contrast, intact<br />

alpha-1 noradrenergic responsivity in OB appears permissive <strong>for</strong><br />

adapting generalization gradients to changes in odor-associated<br />

reward levels. Using computational modeling, we are outlining a<br />

common framework <strong>for</strong> OB processing and neuromodulation to<br />

understand and explain how OB-based circuitry can instantiate<br />

appropriate topologies of learning in response to experience.<br />

Acknowledgements: Supported by NIDCD grant DC009948.<br />

#P93 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

GABAergic gating of olfactory-motor transmission<br />

Gheylen Daghfous 1,2 , Elias Atallah 2 , Jean-Luc Létourneau 1 , François<br />

Auclair 2 , Dominique Derjean 1 , Barbara S Zielinski 3 , Réjean Dubuc 1,2<br />

1<br />

Groupe de Recherche en Activité Physique Adaptée (GRAPA),<br />

Département de kinanthropologie, Université du Québec à Montréal<br />

Montreal, QC, Canada, 2 Groupe de Recherche sur le Système Nerveux<br />

Central (GRSNC), Département de physiologie, Université de Montréal<br />

Montreal, QC, Canada, 3 Department of Biological <strong>Sciences</strong>, University<br />

of Windsor Windsor, ON, Canada<br />

Olfactory stimuli induce and modulate locomotor activity<br />

during vital behaviors such as homing, predator avoidance,<br />

reproduction, <strong>for</strong>aging and feeding. The neural substrate<br />

underlying olfactory-motor behaviors was recently uncovered in<br />

sea lampreys (Petromyzon marinus) [Derjean et al. 2010 PLoS<br />

Biol 8(12): e1000567]. It consists of a specific neural pathway,<br />

extending from the medial part of the olfactory bulb (OB) to<br />

the mesencephalic locomotor region (MLR), with a single relay<br />

in the posterior tuberculum (PT). In all vertebrates, the MLR<br />

acts as a motor command center that controls locomotion<br />

via a descending projection to reticulospinal neurons (RS).<br />

This oligosynaptic pathway permits movements to be rapidly<br />

generated in response to olfactory stimuli, and thus functions<br />

as a pathway dedicated to action. The modulatory mechanisms<br />

that act on this pathway and that are responsible <strong>for</strong> affecting<br />

the variability of the lamprey’s behavioral responses to<br />

olfactory inputs are still unknown. We addressed this question<br />

by using anatomical (tracers and immunohistochemistry)<br />

and physiological (intracellular recordings) techniques.<br />

Retrograde axonal tracing from the PT combined with GABA<br />

immunofluorescence showed dense GABAergic innervation of<br />

the medial OB, a central component of the olfactory-locomotor<br />

pathway, suggesting a role <strong>for</strong> GABA in the modulation of this<br />

pathway. Physiological experiments showed that injections of<br />

the GABA A<br />

antagonist gabazine (0.1 - 1 mM) into the medial<br />

OB considerably amplify or unmask the responses of RS<br />

neurons to olfactory inputs. Taken together, our results suggest<br />

that GABAergic innervation of the OB acts as a gatekeeper <strong>for</strong><br />

sensory inputs to motor control centers. Acknowledgements:<br />

Great Lakes Fishery Commission CIHR NSERC FRSQ<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

64


#P94 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Distinct roles of bulbar muscarinic and nicotinic receptors in<br />

olfactory discrimination learning<br />

Sasha Devore, Licurgo de Almeida, Christiane Linster<br />

Cornell University Department of Neurobiology and Behavior Ithaca,<br />

NY, USA<br />

#P95 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Retronasal odor intensity coding in the dorsal olfactory<br />

bulb of rats<br />

Shree Hari Gautam 1,2 , Michelle R Rebello 1 , Justus V Verhagen 1<br />

1<br />

John B Pierce Lab New Haven, CT, USA, 2 University of Arkansas<br />

Fayetteville, AR, USA<br />

The olfactory bulb (OB) and piri<strong>for</strong>m cortex (PC) receive<br />

dense cholinergic projections from the diagonal band of Broca<br />

in the basal <strong>for</strong>ebrain. Cholinergic modulation within the<br />

PC has long been proposed to serve an important function<br />

in olfactory learning and memory. We here investigate how<br />

olfactory discrimination learning is regulated by cholinergic<br />

modulation of the OB inputs to the PC. Using pharmacological<br />

manipulation of the OB, we examined the role of bulbar<br />

cholinergic signaling in rats’ per<strong>for</strong>mance on a two-alternative<br />

choice odor discrimination task. Results show that blocking<br />

bulbar cholinergic signaling significantly slows learning,<br />

although the relative contribution of muscarinic (MAChRs)<br />

and nicotinic receptors (NAChRs) depends on task difficulty.<br />

Specifically, blocking MAChRs (38 mM scopolamine) impaired<br />

learning <strong>for</strong> nearly all odor sets tested (n=13), whereas blocking<br />

NAChRs (19 mM MLA) only affected learning when the task<br />

was made difficult by using perceptually similar odors. This<br />

pattern of behavioral effects is consistent with predictions from<br />

a recently developed model of cholinergic modulation in the<br />

OB and PC (de Almeida et al., 2012). The model suggests that<br />

MAChRs and NAChRs serve complementary roles in regulating<br />

OB output and cortical learning. Namely, NAChRs determine<br />

the output rate within each OB channel and there<strong>for</strong>e regulate<br />

the overlap between learned representations in the cortical<br />

network. On the other hand, MAChRs control the timing of<br />

spikes across OB output channels and, as a consequence, regulate<br />

the strength of odor representations in the cortical network.<br />

Together, these results suggest that MAChRs in the OB serve a<br />

general role in regulating learning, whereas NAChRs are only<br />

critical when there is substantial overlap in the sensory inputs.<br />

Acknowledgements: NIH R01 DC009948 (CL) NIH F32<br />

DC011974 (SD) L’Oreal Fellowship <strong>for</strong> Women in Science (SD)<br />

In nature food contains many volatile chemicals with a<br />

wide range of concentrations. The volatiles, when released<br />

in the mouth while eating, travel to the nasal cavity via the<br />

nasopharynx evoking a retronasal smell which contributes to<br />

food flavor. The olfactory system is responsible <strong>for</strong> encoding not<br />

only the quality but also the concentration of the volatiles present<br />

in food. It is believed that each odor is represented by a unique<br />

glomerular activation pattern in the olfactory bulb. However,<br />

whether and how retronasal odor concentration is encoded<br />

by the spatiotemporal activity pattern of olfactory glomeruli,<br />

without confounding the quality of a different odorant, remains<br />

unknown. In this study we optically imaged the retronasal<br />

odor-induced calcium responses of olfactory receptor neurons<br />

in the dorsal olfactory bulb in double-tracheotomized rats.<br />

We found reliable concentration-response curves that differed<br />

between odors. MDS of the spatial OB patterns suggest that<br />

ambiguity among select stimuli may occur. Further, the relation<br />

between dynamics and concentration differed remarkably among<br />

retronasal odorants. Understanding of coding <strong>for</strong> retronasal odor<br />

intensity has potentially important implications in the feeding<br />

behavior and flavor neuroscience.<br />

#P96 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Intrinsic oscillatory discharge patterns in mitral cells of the<br />

mouse accessory olfactory bulb<br />

Monika Gorin, Marc Spehr<br />

Dept. of Chemosensation, Institute of Biology II, RWTH Aachen<br />

University Aachen, Germany<br />

The accessory olfactory bulb (AOB) represents the first stage<br />

of central in<strong>for</strong>mation processing in the rodent accessory<br />

olfactory system. In the vomeronasal organ, social chemosignals<br />

activate sensory neurons which <strong>for</strong>m synaptic contacts with<br />

mitral/tufted cells, the main excitatory projection neurons<br />

in AOB. Bypassing the thalamo-cortical axis, these neurons<br />

project directly to higher brain regions such as amygdala<br />

and hypothalamus. Despite their physiological significance,<br />

the intrinsic properties of mitral cells and their role in social<br />

in<strong>for</strong>mation coding and signal integration in the AOB are not<br />

fully understood. Here, we investigate the biophysical properties<br />

of AOB mitral cells using both voltage- and current-clamp whole<br />

cell recordings from optically identified neurons in acute mouse<br />

AOB tissue slices. We identify a population of mitral cells that<br />

display slow oscillatory discharge patterns which persist after<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

65


pharmacological inhibition of synaptic transmission<br />

(AP5, NBQX and gabazine), revealing the network-independent<br />

origin of this bursting behavior. The underlying subthreshold<br />

membrane potential fluctuations with alternating up/down states<br />

display a high degree of periodicity. Using electrophysiological<br />

and pharmacological approaches, we analyze the basic ionic<br />

mechanisms underlying mitral cell oscillatory discharge. Our<br />

data demonstrate a complex interplay of multiple voltage-gated<br />

ionic conductances, such as TTX-sensitive sodium channels<br />

and TEA-sensitive potassium channels as well as conductances<br />

dependent on both intra and extracellular calcium, which<br />

maintains rhythmicity. The oscillatory discharge patterns<br />

observed in AOB mitral cells could play an important role in<br />

the signal coding and / or hormonal homeostasis controlled by<br />

mitral cells target regions in higher brain centers.<br />

#P97 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Lateralized differences in olfactory bulb volume relate to<br />

lateralized differences in olfactory function<br />

Thomas Hummel 1 , Antje Haehner 1 , Cornelia B Hummel 1 , Ilona Croy 2 ,<br />

Emilia Iannilli 1<br />

1<br />

Smell & Taste Clinic, Dept. of ORL, Technical Univ of Dresden<br />

Dresden, Germany, 2 Dept. of Psychosomatic Medicine, Technical Univ<br />

of Dresden Dreden, Germany<br />

The present study aimed to investigate whether side differences<br />

in olfactory bulb (OB) volume correlate to respective differences<br />

in olfactory function. In a total of 164 healthy volunteers<br />

volumetric measures of the OBs were per<strong>for</strong>med plus lateralized<br />

measurements of odor thresholds and odor discrimination. Side<br />

differences were defined as 10% difference between the left and<br />

right OB. In 39 cases volumes on the right side were larger than<br />

on the left side, whereas in 29 cases it was the other way around.<br />

Subjects with larger right-sided OB volumes were found to be<br />

more sensitive to odorous stimulation of the right as compared<br />

to the left nostril in terms of odor thresholds and odor detection;<br />

while correspondingly, higher sensitivity of left nostrils was<br />

observed in individuals with larger OB volumes on the left side.<br />

These data appear to suggest that OB volume is partly dependent<br />

on lateralized influences on the olfactory system, reflecting its<br />

lateralized organization. Acknowledgements: Supported by a<br />

grant from the “Roland Ernst Stiftung” to TH.<br />

#P98 PPOSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Olfactory Sensory Neuron Physiology and Exposure-Induced<br />

Plasticity Are Altered in Adult Olfactory Marker Protein<br />

Knockout Mice<br />

Marley D. Kass, Andrew H. Moberly, John P. McGann<br />

Rutgers University/Psychology Department New Brunswick, NJ, USA<br />

Olfactory marker protein (OMP) is highly and selectively<br />

expressed in olfactory sensory neurons (OSNs) across species,<br />

but its function remains elusive. Previous in vitro studies of<br />

MOR23-expressing OSNs suggested that OMP accelerates the<br />

OSNs’ response to odorants and may modulate the odorantselectivity<br />

of OSNs (Lee et al. 2011 J Neurosci 31:2974-82). Here<br />

we per<strong>for</strong>med in vivo optical imaging in adult mice expressing<br />

the fluorescent exocytosis indicator synaptopHluorin from<br />

the OMP locus. We compared the spatiotemporal patterns of<br />

odor-evoked transmitter release from OSNs in mice that were<br />

heterozygous <strong>for</strong> OMP (OMP-/+) or OMP-null (OMP-/-) and<br />

found that these patterns developed on a slower timescale in<br />

OMP-/- mice but eventually reached the same magnitude as<br />

in OMP-/+ mice. In OMP-/- mice, OSNs innervating a given<br />

glomerulus also responded to a broader range of odorants than<br />

in OMP-/+ mice. These results extend the previous in vitro<br />

findings in MOR23-expressing OSNs to other OSN populations<br />

in vivo. We next evaluated the effects of a 7 day odor exposure<br />

paradigm on these spatiotemporal patterns in adult OMP-/+<br />

and OMP-/- mice. We found that in OMP-/+ mice, odorant<br />

exposure reduced the number of glomeruli receiving OSN<br />

input evoked by the exposure odorant and the magnitude of<br />

those inputs but had no effect on the response to unrelated<br />

odorants. In contrast, in OMP-/- mice this experience-dependent<br />

suppression was observed in the responses to all test odorants,<br />

not just the exposure odorant. These results suggest that OMP<br />

not only conveys odor-selectivity on OSNs but also plays a role<br />

in restricting experience-dependent plasticity to specific OSN<br />

populations. Acknowledgements: This work was supported by<br />

the National Institute on Deafness and Other Communication<br />

Disorders (R00 DC009442 to JPM).<br />

#P99 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Floral Preference is reflected in the Neuroanatomy of the<br />

Olfactory System in Mason Bees (Osmia)<br />

Christina Kelber 1 , Thomas Schmitt 1 , Wolfgang Roessler 2<br />

1<br />

Department of Biology, Ecological Networks TU Darmstadt, Germany,<br />

2<br />

Department of Behavioral Physiology and Sociobiology, Biozentrum<br />

University of Wuerzburg, Germany<br />

POSTER PRESENTATIONS<br />

Locating food sources is important <strong>for</strong> all insects, and in many<br />

species the olfactory system is crucial <strong>for</strong> finding suitable<br />

sources. Two main strategies can be found: either to be able to<br />

use many different food sources (generalists) or to be specialized<br />

on a single or only few food sources (specialists). In bees, both<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

66


generalist and specialized species can be found that collect<br />

pollen and nectar either from many plant families (polylectic)<br />

or from only few species (oligolectic). In some cases, both types<br />

of food preferences can be found within the same genus, like in<br />

the mason bee genus Osmia. Here we investigate how the floral<br />

preference is reflected in the neuroanatomy of the olfactory<br />

system. We employed confocal microscopy scanning and<br />

3D-reconstruction <strong>for</strong> quantitative analyses of major neuropile<br />

volumes. We counted the number of functional units (glomeruli)<br />

within the antennal lobe, the first olfactory neuropile in insects,<br />

and quantified synaptic structures in higher-order sensory<br />

integration centers (mushroom bodies). The investigated Osmia<br />

species showed significant differences in selected neuropile<br />

volumes and also a large interspecies variance in glomerular<br />

numbers, correlated to floral preference. The strictly oligolectic<br />

species Osmia adunca showed the smallest number of glomeruli,<br />

whereas all polylectic species showed larger glomerular numbers.<br />

The mushroom bodies of polylectic and oligolectic species<br />

showed the same density of synaptic structures, but expressed<br />

significant volume differences in the subregions that process<br />

olfactory in<strong>for</strong>mation. Chemical analyzes of host-plant odors<br />

and behavioral tests will be next steps to understand the large<br />

impact of floral preference on the complexity of the olfactory<br />

system in bees. Acknowledgements: DFG KE-1701 1/1<br />

#P100 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Temporal-spatial trans<strong>for</strong>mation in the piri<strong>for</strong>m cortex<br />

Alex Koulakov 1 , Honi Sanders 2 , Brian Kolterman 1 , Dima Rinberg 3 ,<br />

John Lisman 1<br />

1<br />

Cold Spring Harbor Laboratory Cold Spring Harbor, NY, USA,<br />

2<br />

Brandeis University Waltham, MA, USA, 3 New York University<br />

New York, NY, USA<br />

Mitral cells of the olfactory bulb respond to stimuli with<br />

brief and temporally precise transient changes in the firing<br />

rate (sharp events) that tile the inhalation cycle. This suggests<br />

that in<strong>for</strong>mation about odorants can be encoded by the<br />

temporal sequence of these events. Here we propose a class of<br />

computational models <strong>for</strong> the olfactory cortex that can detect<br />

such sequences and convert them into a spatial pattern that<br />

can be recognized by standard attractor networks. We propose<br />

that the olfactory cortex contains groups of cells that can be<br />

sequentially activated by inputs from mitral cells synchronized at<br />

different phases of the respiratory cycle. Neurons in each group<br />

can be persistently activated by virtue of, <strong>for</strong> example, an intrinsic<br />

bistability mechanism. The pattern of activation of neurons<br />

in each group carries a snapshot of coincidences in mitral cell<br />

sharp events at a particular phase of the breathing cycle. Due to<br />

long-range intracortical connectivity, the activation of one group<br />

“enables” bistability in another group which can then <strong>for</strong>m a<br />

snapshot of mitral cell activity at a later phase of the respiratory<br />

cycle. In this way, persistent activation of groups of neuron<br />

occurs sequentially, each in turn representing the olfactory bulb<br />

activity at a certain phase of the sequence. We further show that<br />

sharp events in mitral cell responses occur at a preferred phase of<br />

gamma cycles (measured in the field potential). Given that there<br />

are only a few gamma cycles within a sniff, the number of groups<br />

needed to define gamma cycle specific snapshots of an odorant<br />

is not large. Recognition may occur when the spatial pattern<br />

becomes sufficient to distinguish among the potential odorants.<br />

#P101 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Unique Cholinergic Interneuron Populations in the Mouse<br />

Accessory Olfactory Bulb: Neurochemical Expression and<br />

Fiber Density<br />

Kurt Krosnowski, Sarah Ashby, Weihong Lin<br />

University of Maryland Baltimore County Baltimore, MD, USA<br />

The accessory olfactory bulb (AOB) is a primary central<br />

processing site of sensory in<strong>for</strong>mation detected via the<br />

vomeronasal organ. The AOB contains diverse populations of<br />

intrinsic interneurons. We detected a largely unidentified choline<br />

acetyltransferase-expressing (ChAT) cholinergic interneuron<br />

population using ChAT (BAC) -eGFP mice. Here we classified<br />

their neurochemical expression and distribution throughout<br />

the AOB. We then determined if this cholinergic interneuron<br />

population differs from other known populations of interneurons<br />

in the AOB and main olfactory bulb (MOB). Similar to the<br />

MOB (Krosnowski et al 2012), we found that all cholinergic<br />

interneurons are neither dopaminergic nor GABAergic. While<br />

most ChAT expressing cells in the external plexi<strong>for</strong>m layer (EPL)<br />

of the AOB are not glutamatergic, we found some coexpression<br />

between ChAT-GFP and GluR2/3, a glutamatergic marker, in<br />

contrast with results obtained from the MOB. Also, unlike the<br />

cholinergic interneuron population in the MOB, the majority of<br />

cholinergic interneurons in the AOB do not express a calcium<br />

binding protein, calbindin-D28K. Further, clear differences can<br />

be seen between cholinergic nerve fibers in the internal plexi<strong>for</strong>m<br />

layer (IPL) of the MOB and the AOB. Unlike in the MOB,<br />

where the highest density of cholinergic nerve fibers was found<br />

in the IPL, in the AOB, the IPL contains the fewest visible fibers.<br />

Instead, the majority of cholinergic fibers in the AOB are found<br />

in the EPL. Thus, our data supports the idea that the intrinsic<br />

cholinergic interneuron populations in the AOB are distinct<br />

from previously identified interneuron populations in both<br />

the MOB and AOB and this suggests that they play a unique<br />

role in signal processing in the accessory olfactory system.<br />

Acknowledgements: NIH/NIDCD 009269, 012831 and ARRA<br />

administrative supplement to WL<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

67


#P102 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Blend Processing by Protocerebral Neurons of Manduca sexta<br />

Hong Lei, Hong-Yan Chiu, John Hildebrand<br />

University of Arizona/Department of Neuroscience Tucson, AZ, USA<br />

The male moths of Manduca sexta are more attracted to a mimic<br />

of its natural female sex pheromones, composing of only two<br />

essential components in a ratio that is found in its natural<br />

pheromones. Deviation from this ratio causes reduced behavior.<br />

The projection neurons innervating the pheromone responsive<br />

region of the male antennal lobe produce maximal synchronized<br />

spiking activity in response to blends consisting of the two<br />

components centering around the natural ratio, leading to a<br />

hypothesis that blend ratios are encoded in neuronal synchrony.<br />

To test this hypothesis, we investigated the physiological and<br />

morphological features of down-stream protocerebral neurons<br />

that were challenged with stimulation of single pheromone<br />

components and their blend of different ratios. We found a small<br />

proportion of protocerebral neurons showing stronger responses<br />

to the blend of natural ratio whereas many other neurons<br />

did not distinguish these blends at all. In a multi-dimensional<br />

analysis, we also found the population response mapped onto<br />

the second principle axis displayed most distinction among the<br />

two pheromone components and their blend, and the distinction<br />

occurred prior to the peak population response - a result<br />

consistent with an earlier observation where neural synchrony in<br />

the antennal lobe tends to maximize be<strong>for</strong>e the firing rate reaches<br />

its peak. Moreover, the response patterns of protocerebral<br />

neurons are very diverse, indicating the complexity of internal<br />

representation of odor stimuli at the level of protocerebrum.<br />

Acknowledgements: This work was supported by NSF grant<br />

DMS-1200004 to HL, NIH grant R01-DC-02751 to JGH<br />

#P103 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Comparison of changes in odor-induced firing of mitral cells<br />

and oscillations in the local field potentials in mice learning<br />

to discriminate odors<br />

Anan Li, Diego Restrepo<br />

Department of Cell and Developmental Biology, Rocky Mountain<br />

Taste and Smell Center and Neuroscience Program Aurora, CO, USA<br />

Odor induced mitral cell firing and changes in local field<br />

potentials (LFPs) are modified as an animal learns to<br />

discriminate between odors. In previous work we reported that<br />

as the animal learns to discriminate between odors in go-no go<br />

odor discrimination tasks synchronized unit firing of mitral cells<br />

develop divergent responses to rewarded and unrewarded odors,<br />

and convey important in<strong>for</strong>mation on odor quality in addition to<br />

odor identity in awake behaving mice (Doucette et. al. Neuron<br />

69, 1176–1187, 2011). LFPs reflect integrated signals from cell<br />

ensembles also show divergent responses. However, how mitral<br />

cell firing and local LFPs are related and more importantly how<br />

these are related on a trial-by-trial basis when the animal makes<br />

mistakes remains to be elucidated. Here our preliminary data<br />

indicate that unit firing and beta oscillations of LFPs (10-35<br />

Hz) show related changes during the learning process of the<br />

go-no go task: at the beginning of the task, there is no or very<br />

weak divergent odor responses <strong>for</strong> both signals, while obvious<br />

and strong divergent responses are found as the mice learn to<br />

discriminate the odor pairs. Acknowledgements: DC00566<br />

and DC04657<br />

#P104 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Identification of Microglia in the Peripheral Deafferentation<br />

Response of the Adult Zebrafish Olfactory Bulb<br />

Amanda K McKenna, Christine A Byrd-Jacobs<br />

Western Michigan University/Biological <strong>Sciences</strong> Kalamazoo, MI, USA<br />

Our lab has been examining the potential role of microglia in<br />

the deafferentation response of the zebrafish olfactory bulb. We<br />

previously used phagocytosis-dependent labeling with DiA to<br />

illustrate the putative microglial response following olfactory<br />

organ ablation. DiA-labeled puncta in the deafferented olfactory<br />

bulb increased dramatically in number and then diminished over<br />

the course of a week. The labeling pattern corresponded directly<br />

to areas of the bulb with damaged axons. In that study, we were<br />

unable to identify the labeled profiles conclusively as microglia.<br />

The current study seeks to confirm both the presence and active<br />

role of microglia in the deafferented zebrafish olfactory bulb<br />

using an antibody to zebrafish microglia (anti-4C4). Zebrafish<br />

were treated either with cautery ablation or Triton X-100<br />

application to the olfactory organ to cause either permanent or<br />

temporary deafferentation of the bulb. We hypothesized that the<br />

pattern of anti-4C4 labeling would mimic the pattern seen with<br />

DiA. We found that the olfactory bulb had an obvious increase<br />

in 4C4-positive microglia 1 day following both permanent and<br />

temporary treatments. These 4C4-positive profiles had primarily<br />

amoeboid morphology; they were found throughout the bulb<br />

layers but were concentrated around the degenerating axons.<br />

Over the next several days, the 4C4-positive microglia appeared<br />

to decrease in number; they also changed to mostly ramified<br />

morphologies. This pattern overlaps with the DiA results but also<br />

appears to show additional microglia not actively phagocytizing<br />

axonal debris. Thus, there is a profound microglial response<br />

immediately after both permanent and temporary deafferentation<br />

in the adult zebrafish olfactory bulb that sharply declines over<br />

the next several days. Acknowledgements: Supported by NIH-<br />

NIDCD #011137 (CBJ)<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

68


#P105 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Influences of lateral amygdala activation on piri<strong>for</strong>m cortical<br />

odor processing<br />

Benjamin Sadrian 1,2 , Donald Wilson 1,2<br />

1<br />

NYU School of Medicine New York, NY, USA, 2 Nathan Kline<br />

Institute Orageburg, NY, USA<br />

Olfactory sensory processing in the piri<strong>for</strong>m cortex requires the<br />

synergy of odorant ligand input, local inhibitory feedback loops,<br />

and interregional modulation, in order to synthesize emotionally<br />

relevant and contextually significant odor percepts. Reciprocal<br />

connectivity between the piri<strong>for</strong>m cortex and higher processing<br />

regions, such as the lateral entorhinal cortex and amygdala,<br />

provide currently understudied routes through which odor<br />

processing in the piri<strong>for</strong>m may be regulated. We have employed<br />

optogenetic techniques to investigate how activation of lateral<br />

amygdala (LA) during odor presentation affects piri<strong>for</strong>m cortical<br />

odor processing. We have per<strong>for</strong>med single unit recordings<br />

of both spontaneous and odor-evoked activity in the anterior<br />

piri<strong>for</strong>m cortex, and summarized a range of LA-influenced<br />

changes in piri<strong>for</strong>m activity. We have also begun using a fear<br />

conditioning model to investigate the influences of emotional<br />

significance on odor processing in the piri<strong>for</strong>m. We aim to<br />

describe how such contextual changes affect the precision of<br />

both cortical odor processing and behavioral odor perception.<br />

Acknowledgements: T32-MH067763 from the NIMH to B.A.S.<br />

and R01-DC003906 from the NIDCD to D.A.W.<br />

#P106 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Wild Scents: comparing the olfactory anatomy of caged and<br />

wild mice<br />

Ernesto Salcedo, Kyle Hanson, Taylor Jonas, Lois Low, Diego Restrepo<br />

University of Colorado School of Medicine Aurora, CO, USA<br />

We have previously detailed the subtle neuroanatomical changes<br />

we found in the glomeruli of olfactory bulbs from genetically<br />

identical mice reared in cages with different levels of ventilation<br />

(Oliva and Salcedo et al, 2010). In these mice, we were able<br />

to correlate these glomerular changes with marked increases<br />

in aggressive behavior towards invader mice, highlighting the<br />

exquisite sensitivity a mouse’s olfactory neuroanatomy has to<br />

its environment. In order to examine the broader effects that<br />

environment may have on the <strong>for</strong>mation of the olfactory system,<br />

we have trapped wild house mice from the Denver environs<br />

and have rigorously characterized the neuroanatomy of their<br />

main olfactory bulbs (MOB) using MATLAB mapping software<br />

developed in-house and immunohistochemical techniques. On<br />

gross examination, the MOBs from the wild mice do not appear<br />

to be significantly different from their caged brethren. Nor did<br />

we find any significant immunostaining differences in OMP<br />

of GAP43 labeling of the MOB. Although somewhat smaller,<br />

the wild olfactory bulbs had an estimated number of glomeruli<br />

(using Meisami’s Correction) that does not differ significantly<br />

from the estimated number of glomeruli found in the MOBs<br />

of their caged counterparts. Curiously, we do find a dramatic<br />

difference in the distribution of olfactory sensory innervation<br />

across the surface of the MOB: caged mice tended to have larger<br />

glomeruli that occupied a significantly larger portion of the<br />

glomerular layer then did the wild mice. This distribution was<br />

particularly pronounced in the ventro-medial portion of the bulb<br />

around the AOB. These results provide further evidence that<br />

olfactory environment plays a role in fine-tuning the <strong>for</strong>mation<br />

and maintenance of glomeruli in the main olfactory bulb.<br />

Acknowledgements: NIDCD<br />

#P107 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Assessment of nasally administered insulin-like growth<br />

factor-I accumulation in the cerebrum of mice with resected<br />

olfactory bulb<br />

Hideaki Shiga 1,2 , Mikiya Nagaoka 2 , Kohshin Washiyama 2 ,<br />

Junpei Yamamoto 1 , Ryohei Amano 2 , Takaki Miwa 1<br />

1<br />

Otorhinolaryngology, Kanazawa Medical University Ishikawa, Japan,<br />

2<br />

Quantum Medical Technology, Kanazawa University Ishikawa, Japan<br />

Objectives: To show the role of the olfactory bulb in the delivery<br />

of nasally administered insulin-like growth factor-I (IGF-I) to the<br />

brain in vivo. Nasal administration of IGF-I has been shown to<br />

enable drug delivery to the brain beyond the blood brain barrier<br />

in vivo. IGF-I is associated with the development and growth<br />

of the central nerve. Methods: The ratio of uptake of nasally<br />

administered 125 I-IGF-I in the cerebrum to uptake in the blood<br />

of male ICR mice with resected left olfactory bulb (8 weeks of<br />

age, the model mice) was compared to that of the sham-operated<br />

male ICR mice (8 weeks of age, the control mice). We exposed<br />

and resected the left olfactory bulb, cutting the frontal bones of<br />

model mice, and just exposed the left olfactory bulb in control<br />

mice under anesthesia. 125 I-IGF-I (human, recombinant) saline<br />

solution was obtained from PerkinElmer Japan (Yokohama,<br />

Japan), and 10μl was instilled into the left nostril of each mouse<br />

with a microinjection pipette under anesthesia. The radioactivity<br />

of the samples was measured with gamma spectrometry. The<br />

accumulation of the nasally administered neuronal tracer<br />

(fluoro-ruby; dextran tetramethylrhodamine) in the epithelium<br />

of mice was assessed in frozen sections under a fluoroscopic<br />

microscope. Results: The ratio of uptake of nasally administered<br />

125<br />

I-IGF-I in the cerebrum to uptake in the blood of the model<br />

group was significantly decreased compared to the control group.<br />

The accumulation of nasally administered neuronal tracer in<br />

the nasal epithelium of mice was significantly prevented by the<br />

resection of the olfactory bulb. Conclusions: Olfactory bulb<br />

resection results in the reduced delivery of nasally administered<br />

IGF-I to the brain due to the disconnection of the olfactory<br />

nerve between the nasal epithelium and olfactory bulb in vivo.<br />

Acknowledgements: Grant-in-Aid <strong>for</strong> Scientific Research<br />

from the Ministry of Education, Science and Culture of Japan<br />

(C21592174 to H.S.) and Assist Kaken from Kanazawa Medical<br />

University (J.Y.)<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

69


#P108 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

#P109 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Dynamics of reward-mediated neural plasticity in honey<br />

bee antennal lobe glomerulus<br />

Adrian Smith 1,4 , Irina Sinakevitch 1 , Ramon Huerta 2 ,<br />

Maxim Bazhenov 3 , Brian H Smith 1<br />

1<br />

Arizona State University, School of Life Science, Tempe, AZ, USA,<br />

2<br />

BioCircuits Institute, University of Cali<strong>for</strong>nia San Diego, La Jolla,<br />

CA, USA, 3 Department of Cell Biology and Neuroscience, University<br />

of Cali<strong>for</strong>nia, Riverside, CA, USA, 4 Arizona State University,<br />

Mathematical, Computational and Modeling <strong>Sciences</strong> Center, Applied<br />

Mathematics <strong>for</strong> the Life and Social <strong>Sciences</strong> Tempe, AZ, USA<br />

Recent studies of neural plasticity in the honey bee antennal lobe<br />

(AL) show the response dynamics of uniglomerular projection<br />

neurons (uPNs) to an odor change after association of that odor<br />

with sucrose rein<strong>for</strong>cement. The octopamine (OA), released by<br />

the ventral unpaired median neuron (VUM), is necessary <strong>for</strong><br />

these neural plasticity changes. Using anti-OA staining, we found<br />

that the varicosity-like distributions of VUM branch mostly in<br />

the cortex of the glomerulus, where it potentially modulates<br />

olfactory receptor neurons (ORNs), local interneurons (LNs)<br />

and uPNs. We develop a biophysical model of the AL circuit<br />

to investigate modulatory mechanisms that can explain existing<br />

data on the dynamical changes of uPNs during associative<br />

learning, and lead to insights <strong>for</strong> new experiments. First, we<br />

hypothesize that OA release from VUM varicosities would be<br />

dependent on correlated firing between ORNs and VUM during<br />

associative learning. This mechanism implies simultaneous<br />

cholinergic and octopaminergic transmission to PNs. Second,<br />

OA release from VUM acts on AmOA1 receptors expressed<br />

in LNs. AmOA1 activation increases the excitability of LNs,<br />

leading to increased inhibition of PNs. Third, release of OA<br />

leads to direct activation of beta-adrenergic-like OA receptors.<br />

This increases the levels of cAMP, triggering PKA-dependent<br />

translation and upregulation of alpha7 nicotinic acetylcholine<br />

receptor (nAChR) subtypes. nAChR-dependent Ca2+ influx<br />

triggers transcription factors that upregulate transient Shal-type<br />

K+ channels, preventing excessive membrane depolarization.<br />

Based on these hypotheses, we propose the model to characterize<br />

dynamics of the uPN as a function of the relative expression<br />

of Shal-type K+ channels and OA release. Acknowledgements:<br />

NIH-NCRR RR014166 and NIH grant R01 DC011422<br />

In vivo imaging of odor-evoked responses in the mouse<br />

olfactory bulb using the FP voltage sensor ArcLight and the<br />

calcium sensor GCaMP3<br />

Douglas A Storace 1 , Lawrence B Cohen 1, 2 , Uhna Sung 2<br />

1<br />

Yale University / Department of Cellular and Molecular Physiology<br />

New Haven, CT, USA, 2 Korea Institute of Science and Technology /<br />

Center <strong>for</strong> Functional Connectomics Seoul, South Korea<br />

Optogenetic reporters of membrane potential allow <strong>for</strong> recording<br />

of genetically distinct populations of neurons, although their<br />

usefulness to date has been limited by poor in vivo expression,<br />

small signal sizes and slow kinetics. The fluorescent protein (FP)<br />

voltage sensor ArcLight exhibits a change in fluorescence to a<br />

100 mV depolarization five times larger than previously reported<br />

probes in HEK 293 cells. However, recordings of ArcLight in<br />

mammalian neurons have been limited to cultured neurons. The<br />

goal of the present study was to examine ArcLight responses in<br />

the olfactory bulb in an in vivo preparation, and compare them<br />

to those of the genetically encoded calcium indicator GCaMP3.<br />

AAV-1 viral transduction was used to express ArcLight and<br />

GCaMP3 in the mouse olfactory bulb. Odors were presented at<br />

different stimulus duration and concentrations, and the resulting<br />

patterns of activation were imaged. Odor-specific patterns of<br />

activation were obtained from both ArcLight and GCaMP3,<br />

although only ArcLight had sufficiently fast temporal kinetics<br />

to clearly detect population activity elicited by individual sniffs<br />

to an odor. The results indicate that ArcLight can be used as a<br />

reliable detector of odor-evoked population signals in the mouse<br />

olfactory bulb. Acknowledgements: Supported by US NIH<br />

Grants DC005259 and NS057631, Grant WCI 2009-003 from<br />

the National Research Foundation of Korea, and an<br />

James Hudson Brown – Alexander Brown Coxe Fellowship<br />

from Yale University.<br />

#P110 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Expression and Activity of Glucagon-like Peptide-1 in the<br />

Mouse Olfactory Bulb<br />

Nicolas Thiebaud 1 , Ida Llewellyn-Smith 2 , Fiona Gribble 3 , Frank<br />

Reimann 3 , Stefan Trapp 4 , Debra A. Fadool 5<br />

1<br />

The Florida State University, Department of Biological Science<br />

Tallahassee, FL, USA, 2 Flinders University, Centre <strong>for</strong> Neuroscience<br />

Bed<strong>for</strong>d Park, Australia, 3 Addenbrooke’s Hospital, Cambridge Institute<br />

<strong>for</strong> Medical Research Cambridge, United Kingdom, 4 Imperial College<br />

London, Department of Surgery and Cancer London, United Kingdom,<br />

5<br />

The Florida State University, Program in Neuroscience and Molecular<br />

Biophysics Tallahassee, FL, USA<br />

POSTER PRESENTATIONS<br />

A number of peptides and hormones that are known to regulate<br />

energy metabolism or feeding behavior have been identified in<br />

the olfactory system. These hormones are thought to modulate<br />

olfactory perception and function to suppress or promote<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

70


appetite. Glucagon-like peptide-1 (GLP-1) is an incretin peptide<br />

that also suppresses food intake, and in situ hybridization in<br />

rat has suggested that both GLP-1 and its receptor are present<br />

in the olfactory bulb (OB). Furthermore, GLP-1 modulation<br />

of taste sensitivity has been reported <strong>for</strong> taste buds, prompting<br />

us to speculate a similar role in the olfactory system. Using<br />

transgenic mice expressing YFP under the preproglucagon (PPG)<br />

promoter, we confirm that a population of YFP-immunoreactive<br />

neurons exists in the granule cell layer (GCL) of the OB,<br />

indicative of GLP-1 producing cells. The labeled neurons had<br />

a typical granule cell morphology with a single axon projecting<br />

into the adjacent mitral cell layer (MCL). We observed strong<br />

immunoreactivity <strong>for</strong> the GLP-1 receptor in the MCL and in<br />

sparse cells in the GCL in olfactory marker protein (OMP-GFP)<br />

transgenic mice. Binding of biotinylated GLP-1 to the GLP-1<br />

receptor was visualized within the same regions. We confirmed<br />

the presence of functional GLP-1 receptors on mitral cells (MCs)<br />

using whole-cell patch-clamp recordings in acute OB slices.<br />

Bath perfusion of 1 µM GLP-1 or its stable analogue, exendin-4,<br />

resulted in a significant increase in the evoked action potential<br />

frequency (370% <strong>for</strong> GLP-1 and 170% <strong>for</strong> exendin-4) and a<br />

concomitant decrease in the interburst interval in about 60% of<br />

the sampled MCs. These results show that GLP-1 is synthesized<br />

locally in the OB and directly affects the firing properties of<br />

MCs, suggesting a potential paracrine modulation of olfactory<br />

output with satiety. Acknowledgements: This work was<br />

supported by R01 DC003387 from the NIH/NIDCD, a Creative<br />

Research Council (CRC) award from FSU, a Project Grant<br />

1025031 from NHMRC Australia, and MR/J013293/1 from the<br />

MRC United Kingdom.<br />

#P111 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Enhanced Survival of Newly Formed Cells Contributes to<br />

Restoration of Olfactory Bulb Volume Following Reversible<br />

Deafferentation in Adult Zebrafish<br />

Darcy M Trimpe, Christine A Byrd-Jacobs<br />

Western Michigan University/Biological <strong>Sciences</strong> Kalamazoo, MI, USA<br />

Our lab has shown that chronic partial deafferentation, achieved<br />

through unilateral chemical ablation of the olfactory epithelium<br />

with Triton X-100, results in a decrease in olfactory bulb<br />

volume, while cessation of treatment allows bulb volume to<br />

recover. We hypothesized that alterations in cell genesis and/<br />

or survival would be involved in restoration of olfactory bulb<br />

size. Bromodeoxyuridine (BrdU) administration was used to<br />

examine newly <strong>for</strong>med cells in the brain of adult zebrafish,<br />

with short-term survival allowing investigation of patterns of<br />

cell proliferation and long-term survival allowing examination<br />

of cell survival and fate. We first compared two methods of<br />

BrdU administration: immersion of fish in the drug versus<br />

intraperitoneal injection. While both methods revealed similar<br />

numbers of newly <strong>for</strong>med cells, injection of the drug resulted<br />

in loss of fewer fish during treatment. Next, we examined<br />

potential alterations in cell genesis and/or cell survival resulting<br />

from reversible partial deafferentation. Repeated detergent<br />

treatment followed by BrdU exposure showed no difference in<br />

the number of dividing cells in the olfactory bulb, indicating that<br />

cell genesis is not affected. There was, however, an increase in<br />

newly <strong>for</strong>med cells that survived when the detergent treatment<br />

ceased, indicating that cell survival contributes to the restoration<br />

of bulb volume during the period of reinnervation. When fish<br />

were exposed to BrdU be<strong>for</strong>e the repeated detergent treatment,<br />

there appeared to be no effect on the number of newly <strong>for</strong>med<br />

cells. Thus, enhanced cell survival, rather than cell genesis,<br />

appears to be a contributing factor in the restoration of olfactory<br />

bulb volume following return of innervation in a reversible<br />

deafferentation model. Acknowledgements: Supported by<br />

NIH-NIDCD #011137 (CBJ)<br />

#P112 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

Characterizing Olfactory Bulb Circuitry using Intrinsic<br />

Flavoprotein Fluorescence Imaging<br />

Cedric R Uytingco 1,2,4 , Adam C Puche 1,2,4 , Steven D Munger 1,2,3,4<br />

1<br />

Depatment of Anaotmy and Neurobiology Baltimore, MD, USA,<br />

2<br />

Program in Neuroscience Baltimore, MD, USA, 3 Department of<br />

Medicine, Division of Endocrinology, Diabetes and Nutrition Baltimore,<br />

MD, USA, 4 University of Maryland School of Medicine Baltimore,<br />

MD, USA<br />

The main olfactory system has distinct subsystems that differ<br />

in the chemosensory stimuli to which they respond and the<br />

connections they make to the brain. In contrast to the canonical<br />

glomeruli of the main olfactory bulb (MOB), individual necklace<br />

glomeruli (NGs) receive heterogeneous olfactory sensory neuron<br />

(OSN) innervation [including semiochemical-sensitive OSNs<br />

expressing guanylyl cyclase D (GC-D)] and display extensive<br />

intrabulbar connections. This organization suggests that NGs<br />

integrate multiple olfactory signals. To better understand the<br />

functional circuitry associated with NGs, we examine the transsynaptic<br />

spread of neuronal activity using intrinsic flavoprotein<br />

fluorescence (FF) imaging. FF imaging measures the changes<br />

in endogenous fluorescence produced by mitochondrial<br />

flavoproteins that accompany increased metabolic demand. The<br />

FF signals correspond to neuronal activity and can be followed<br />

across synapses, thus facilitating the mapping of functional<br />

circuits. Studies in horizontal MOB slices from 3-5 w.o. mice<br />

exhibit robust FF signal spread from the glomerular layer<br />

(GL) to the external plexi<strong>for</strong>m layer (EPL) following electrical<br />

stimulation (1-4 s, 10-50 Hz, 10-100 μA) of individual canonical<br />

and necklace glomeruli. Bath application of 10 mM gabazine<br />

increased lateral stimulus-dependent FF signal spread in the GL/<br />

EPL, and resulted in a 2-fold increase in FF signal amplitude in<br />

the EPL. Single NG stimulation (from mice expressing green<br />

fluorescent protein under control of the gene encoding GC-<br />

D) results in reduced FF signal amplitude but prolonged FA<br />

signal duration compared to canonical glomeruli. The use of<br />

FF imaging should help reveal basic strategies of in<strong>for</strong>mation<br />

processing in the MOB and its subsystems. Acknowledgements:<br />

NIDCD (DC005633), NIGMS (GM008181), NINDS<br />

(NS063391)<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

71


#P113 POSTER SESSION II:<br />

OLFACTION DEVELOPMENT; TASTE CNS;<br />

NEUROIMAGING; OLFACTION CNS<br />

#P114 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

DIFFERENTIAL MODIFICATIONS OF SYNAPTIC<br />

WEIGHTS DURING ODOR RULE LEARNING:<br />

DYNAMICS OF INTERACTION BETWEEN THE<br />

PIRIFORM CORTEX WITH LOWER AND HIGHER<br />

BRAIN AREAS<br />

Yaniv Cohen 1,2,3 , Donald A. Wilson 2,3 , Edi Barkai 1<br />

1<br />

Departments of Biology and Neurobiology, Faculty of Natural <strong>Sciences</strong>,<br />

University of Haifa Haifa, Israel, 2 Department of Child and Adolescent<br />

Psychiatry, New York University Langone School of Medicine New<br />

York, NY, USA, 3 Emotional Brain Institute, Nathan Kline Institute <strong>for</strong><br />

Psychiatric Research, Orangeburg New York, NY, USA<br />

Learning of a particularly difficult olfactory-discrimination (OD)<br />

task results in acquisition of rule-learning, a process that requires<br />

prolonged and extensive training. Previously, we demonstrated<br />

enhanced synaptic connectivity between the piri<strong>for</strong>m cortex (PC)<br />

and its ascending and descending inputs from the olfactory bulb<br />

(OB) and orbitofrontal cortex (OFC) following OD rule learning.<br />

Here, using recordings of evoked field post-synaptic potentials in<br />

behaving animals, we examined the dynamics by which synaptic<br />

connectivity from the OB and OFC to the PC are modified<br />

during rule acquisition. We show profound differences in the<br />

dynamics and strength of synaptic connectivity modulation<br />

between the ascending and descending inputs. During rule<br />

learning acquisition, the ascending synaptic connectivity from<br />

the OB to the anterior and posterior PC is simultaneously<br />

enhanced. Notably, the daily OB electrical stimulation used to<br />

examine the strength of synaptic inputs enhanced the rate of rule<br />

learning. In sharp contrast, the synaptic input in the descending<br />

pathway from the OFC was significantly reduced during rule<br />

learning acquisition. OFC stimulation had no effect on the rate at<br />

which the rule was acquired. Once rule learning was established,<br />

the strength of synaptic connectivity in the two pathways<br />

resumed its pre-training values. We suggest that acquisition<br />

of olfactory rule learning requires a transient enhancement of<br />

ascending inputs to the PC, synchronized with a parallel decrease<br />

in the descending inputs. This combined short-lived modulation<br />

is required to enable the PC network to reorganize in a manner<br />

that enables it to first acquire and then maintain the rule.<br />

Human exposure to acrolein – time dependence on<br />

TRPA1 agonists<br />

Anna-Sara Claeson, Nina Lind<br />

Department od Psychology, Umeå university Umeå, Sweden<br />

The objective of the study was to examine the time dependence<br />

on sensory irritation potency of acrolein (2-propenal) in humans.<br />

Concentrations at or below earlier reported thresholds that<br />

initially are too low to evoke sensory irritation in the eye but<br />

might do so in exposures up to 60 minutes were used. Acrolein<br />

is a known TRPA1 agonist present in cigarette smoke, smoke<br />

from fires, automobile exhaust and smog. The TRPA1 channel<br />

is activated by electrophilic compounds that <strong>for</strong>m covalent<br />

bonds with cysteine residues. Because of this mechanism<br />

of activation one can expect duration of exposure to be of<br />

importance in evoking sensory irritation. The exposures occurred<br />

in an exposure chamber and the subjects were breathing fresh<br />

air through a mask that covered the nose and mouth. All<br />

participants took part in four exposure conditions, differing in<br />

duration and concentration. The concentrations of acrolein<br />

(diluted in heptane) were 0.35, 0.07, 0.05 and 0 ppm (during<br />

15, 45, 60 and 30 min, respectively). During the 30 minutes of<br />

exposure subjects were exposed to only heptane at the same<br />

concentration as in the other exposures (4.9 ppm). During<br />

exposure, eye irritation was rated on Borg’s CR-100 scale.<br />

Human exposure to acrolein at sub-threshold concentrations<br />

showed a cumulative effect on sensory irritation. During<br />

exposure to 0.35 ppm (but not 0.07 and 0.05 ppm) acrolein<br />

evoked a significant increase in irritation compared to the control<br />

condition after about 12 minutes of exposure. During exposure<br />

to 0.07 and 0.05 ppm only some of the subjects reported<br />

increased irritation after about 30 minutes of exposure. A large<br />

variability in reported sensory irritation was seen between<br />

individuals and this may be due to individual differences in the<br />

ability to remove the electrophilic irritants from the cornea.<br />

Acknowledgements: The Swedish Research Council FORMAS<br />

#P115 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Solitary Chemosensory Cells in Human Nasal Epithelium<br />

Sarah E Cooper 1 , Marco Tizzano 2 , Vijay R Ramakrishnan 1 ,<br />

Henry P Barham 1 , Jameson K Mattingly 1 , Thomas E Finger 1,2 ,<br />

Sue C Kinnamon 1,2<br />

1<br />

University of Colorado, Department of Otolaryngology Aurora, CO,<br />

USA, 2 University of Colorado, Department of Cell and Developmental<br />

Biology Aurora, CO, USA<br />

POSTER PRESENTATIONS<br />

Solitary chemosensory cells (SCCs) described in rodents rely on<br />

the bitter taste transduction cascade to detect potential irritants<br />

within the airways. SCCs express all of the taste GPCR signaling<br />

effectors including T2R bitter taste receptors, a-gustducin,<br />

PLCb2, and the transduction channel TRPM5. SCCs are<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

72


innervated by the trigeminal nerve and when stimulated evoke<br />

protective airway reflexes such as sneezing, apnea and local<br />

inflammation (Tizzano et al., 2010; AChemS 2012). We have<br />

begun investigating SCCs in human sinonasal epithelium and<br />

their clinical implications. Previously we and others reported<br />

that cells resembling SCCs occur in human biopsy material<br />

near the vestige of the vomeronasal organ (Braun et al., 2011)<br />

as well as within the turbinates (Barham et al 2013; AChemS<br />

2012). However, the exact distribution and abundance of SCCs<br />

in humans is unknown. To map the distribution of SCCs, we<br />

obtained middle and inferior turbinate mucosa from human<br />

patients that were free of sinonasal disease, but were undergoing<br />

surgical procedures requiring removal of this tissue. Whole<br />

mount tissue was stained with antibodies against TRPM5 and<br />

villin, which is expressed at the apex of microvillous, but not<br />

ciliated, epithelial cells. TRPM5-immunoreactive cells were<br />

scattered heterogeneously in the sinonasal tissue. The cells were<br />

most abundant on the ridges of the turbinates and less abundant<br />

on the lateral margins. Many TRPM5 immunoreactive cells<br />

also labeled with the villin antibody, suggesting that TRPM5 is<br />

present in microvillous but not ciliated cells of the epithelium.<br />

Studies are in progress to determine if disease state alters the<br />

distribution or abundance of these cells and whether SCCs in<br />

humans are innervated by the trigeminal nerve, as in rodents.<br />

Acknowledgements: R01 DC009820 (TEF and SCK) P30<br />

DC04657 (to D. Restrepo)<br />

#P116 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Responses to change in oral temperature by neurons in the<br />

mouse medullary dorsal horn and nucleus of the solitary tract<br />

Yi Kang, Christian Lemon<br />

St. Louis University St. Louis, MO, USA<br />

Psychophysical data show the temperature of sapid solutions<br />

influences flavor. However, it is not entirely clear how central<br />

neural circuits <strong>for</strong> oral sensation encode temperature input.<br />

The trigeminal subnucleus caudalis (Vc) is a brainstem somatic<br />

relay receiving temperature signals from the oro-facial region<br />

and implicated <strong>for</strong> oral thermosensation. Additionally, the<br />

solitary tract nucleus (NTS), the first central taste relay, also<br />

receives thermal input from the mouth. Here we compared<br />

neural responses of NTS and Vc to intraoral thermal stimulation<br />

in anesthetized mice to assess the contribution of activity in<br />

these structures to oral temperature responses in brain stem.<br />

Extracellular single-unit activities of NTS thermo-gustatory<br />

neurons and Vc thermo-somatic cells were recorded after<br />

application of different temperature stimuli to tongue, including<br />

cold (5 and 10 °C), cool/ambient (17 and 23 °C), and warm/<br />

hot (30, 45, and 48 °C). Temperature stimulation was achieved<br />

rapidly by oral flow of thermally varied water. Seventytwo<br />

neurons were obtained; 34 from Vc and 38 from NTS.<br />

Analyses revealed significant differences between NTS and<br />

Vc in responses to thermal stimuli [F(1, 70) = 5.80, P ATP > acetylcholine<br />

(ACh)). Because of the low percentage of ACh-sensitive MCs<br />

and SCCs, ACh released from these cells may play a role in<br />

paracrine regulation to influence neighboring cells. Using Ca 2+<br />

imaging on intact epithelial preparations, we found that AChinduced<br />

increases in intracellular Ca 2+ levels in epithelial cells<br />

surrounding the MCs and SCCs were inhibited by muscarinic<br />

ACh antagonist atropine. Our results suggest that MCs and<br />

SCCs share common physiological roles in sensing chemical<br />

stimuli and may release ACh to influence surrounding nonsensory<br />

cells via paracrine mechanism. Because MCs lack<br />

afferent innervation, this cholinergic paracrine regulation could<br />

be especially important <strong>for</strong> chemoreception-mediated regulation<br />

of MOE activities. Acknowledgements: Supported by research<br />

grants NIH/NIDCD 009269, 012831 and ARRA administrative<br />

supplement to WL<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

73


#P118 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Long-term acclimation to capsaicin solutions affects taste bud<br />

volume and consumption in rats<br />

Jacquelyn M Omelian, Suzanne I Sollars<br />

University of Nebraska at Omaha/Psychology Omaha, NE, USA<br />

The effects of chronic exposure to capsaicin (the component<br />

responsible <strong>for</strong> the piquancy of chili peppers) in the gustatory<br />

system are not yet well understood; a critical consideration given<br />

capsaicin’s popularity in culinary and medicinal applications.<br />

To examine these potential effects, rats received 40 days of<br />

treatment with a 30% sucrose solution containing either a 5 ppm<br />

capsaicin (in 2.5% ethanol) or sham (2.5% ethanol) condition.<br />

Animals began exposure as neonates (P5) or adults (P40) to<br />

evaluate potential differences across development. Taste bud<br />

volumes within fungi<strong>for</strong>m papillae were measured at either two<br />

or 50 days post treatment, to assess immediate or lasting effects.<br />

Animals treated with capsaicin as neonates had significantly<br />

smaller taste buds at 50 days post treatment but in no other group<br />

(neonate or adult) were there significant differences. Capsaicin<br />

is a trigeminal irritant and does not directly affect the chorda<br />

tympani-innervated taste buds, thus the difference in taste bud<br />

volumes following capsaicin exposure suggests an integrated<br />

relationship between the chorda tympani and lingual nerves in<br />

gustatory maintenance. The capsaicin concentrations used here<br />

were significantly lower than those found in nature, as treatments<br />

were limited to what animals would consume willingly. As an<br />

additional experiment, we examined whether rats develop a<br />

tolerance to capsaicin over time. To do so we gave adult animals<br />

5 ppm capsaicin/30% sucrose solutions and then incrementally<br />

increased the quantity of capsaicin by 2.5 ppm after each 5 day<br />

period. Initial results showed animals willingly consumed higher<br />

levels of capsaicin (10 ppm) with this type of acclimation. Taste<br />

bud volume analyses <strong>for</strong> these animals will be presented, and<br />

further experiments with additional acclimation are ongoing.<br />

Acknowledgements: University of Nebraska at Omaha:<br />

Graduate Research and Creative Activities Fund<br />

#P119 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Trigeminal Mediation of Mammalian Aversion to Insect<br />

Chemical Defense Compounds<br />

Paige Richards, Pamela Fazio, Alexa Ciesinski, Annalyn Welp,<br />

Deirdre Craven, Wayne Silver<br />

Wake Forest University Winston-Salem, NC, USA<br />

Insects release defensive chemicals as one mechanism <strong>for</strong><br />

deterring predation and defending their territory. A wide<br />

variety of insects use the same chemicals <strong>for</strong> this purpose. We<br />

hypothesize that defensive insect chemicals used by multiple<br />

insect orders are targeting the mammalian trigeminal system and<br />

eliciting chemesthesis. In the current study, we test a number of<br />

insect defense compounds to determine if they are irritating to a<br />

mammalian predator (rat). We first determined if these defense<br />

compounds activated the trigeminal nerve by recording from<br />

the ethmoid branch and monitoring respiration while perfusing<br />

stimuli through the rat’s nasal cavity. Using these methods, we<br />

determined that all of the tested compounds except tetradecane<br />

activate the rat trigeminal nerve. Then, we examined behavioral<br />

aversion responses to these compounds by placing a rat in a<br />

square plexiglass arena with petri dishes in each corner. After<br />

a period of habituation, an insect chemical defense compound<br />

was added to one of the four petri dishes and the movement of<br />

the rat recorded. Ethovision XT (Noldus) was used to analyze<br />

whether rats spent less time in the irritant containing corner. Rats<br />

were behaviorally averse to most compounds at concentrations<br />

similar to those released by insects. By imaging primary cultures<br />

of rat trigeminal ganglia using a calcium responsive dye, we<br />

have confirmed the activation of the trigeminal system by<br />

insect chemical defense compounds. Additional experiments<br />

are underway to determine the specific receptor targets of these<br />

chemicals using a heterologous expression system. In the future,<br />

we will be doing similar work with chickens to determine the<br />

role of the trigeminal system in avian responses to insect defense<br />

compounds. Acknowledgements: WFU Center <strong>for</strong> Molecular<br />

Communication and Signaling<br />

#P120 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Cholingeric Neurotransmission Links Solitary Chemosensory<br />

Cells To Nasal Inflammation<br />

CJ Saunders 1,2 , Thomas E Finger 1,2 , Marco Tizzano 1<br />

1<br />

Rocky Mtn Taste & Smell Center, Univ Colo School of Medicine<br />

Aurora, CO, USA, 2 Neuroscience Program, Univ Colo School of<br />

Medicine Aurora, CO, USA<br />

Solitary chemosensory cells (SCCs) are specialized epithelial<br />

chemosensors that respond to “bitter” substances via the<br />

canonical taste transduction cascade (T2Rs, Ga-gustducin<br />

and TRPM5). When stimulated, SCCs release a hitherto<br />

unidentified neurotransmitter onto peptidergic nociceptive<br />

trigeminal fibers. Activation of these nociceptive fibers triggers<br />

neurogenic inflammation via NK1 (substance P) receptors on<br />

capillaries causing local plasma extravasation (Tizzano & Finger,<br />

AChemS 2012). How SCCs activate nerve fibers is unknown.<br />

In the present study, we show that choline acetyltransferase,<br />

the synthetic enzyme <strong>for</strong> acetylcholine (ACh), is present in<br />

SCCs. Previous studies have shown nicotinic ACh receptors<br />

(nAChR) on trigeminal fibers. Thus, all elements <strong>for</strong> cholinergic<br />

neurotransmission are present in the SCC-trigeminal system.<br />

To test if SCC-mediated inflammation requires activation of<br />

nAChRs, we measured SCC-evoked plasma extravasation in<br />

mice stimulated unilaterally with denatonium benzoate (20 μL,<br />

10mM). Prior to chemical stimulation, mice were injected i.p.<br />

with either saline or the nAChR-antagonist Mecamylamine<br />

(Mec). Under urethane anesthesia, the right naris was stimulated<br />

with denatonium, and the mouse injected i.v. with Alexa555-<br />

conjugated albumin. Heads were bisected and fluorescence<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

74


intensity in the nasal epithelium was quantified. A one-way<br />

ANOVA demonstrated a significant difference between the three<br />

experimental groups [F(2,12)=17.03; p


nerve fibers. Nasal SCCs respond to bitter compounds including<br />

bacterially-produced molecules, to evoke protective respiratory<br />

reflexes and early inflammatory responses (Gulbransen 2008,<br />

Tizzano 2010 and AChemS 2012). Here we test whether SCCmediated<br />

pro-inflammatory responses require activation of the<br />

trigeminal nerve and subsequent peptidergic neurotransmission,<br />

or whether SCC activation triggers local inflammation via<br />

an intramucosal paracrine signaling mechanism. When<br />

denatonium (10mM) is instilled into the nasal passageways,<br />

it evokes SCC-dependent plasma leakage and local mast cell<br />

(MC) degranulation (Tizzano AChemS 2012). Chemical<br />

ablation of peptidergic nerve fibers with resiniferatoxin (RTX,<br />

an ultrapotent analog of capsaicin) eliminates both denatoniummediated<br />

plasma leakage and MC degranulation. These results<br />

show that the peptidergic nerve fibers are necessary <strong>for</strong> these<br />

SCC-mediated pro-inflammatory events. Moreover, injection of<br />

L732138 (5mg/kg), an inhibitor of the neurokinin 1 (substance<br />

P) receptor present on blood vessels, prevents denatoniuminduced<br />

plasma leakage. This indicates that substance P is the<br />

mediator <strong>for</strong> plasma leakage. Our results demonstrate that<br />

activation of the SCCs leads to a rapid, local pro-inflammatory<br />

response via neurogenic mechanisms. This fast pro-inflammatory<br />

response, driven by the SCCs in conjunction with the trigeminal<br />

nerve, represents a 1st line of defense against respiratory<br />

epithelial assault by noxious chemicals and bacterial pathogens.<br />

Acknowledgements: NIDCD R03 DC012413 (M.T.), R01<br />

DC009820 (T.E.F.), and P30 DC04657 (to D. Restrepo)<br />

#P124 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

The Influence of Bubbles on the Perception of<br />

Carbonation Bite<br />

Paul M Wise 1 , Stephen R Thom 2 , Madeline Wolf 1 , Bruce Bryant 1<br />

1<br />

Monell Chemical Senses Center Philadelphia, PA, USA,<br />

2<br />

University of Pennsylvania/Institute <strong>for</strong> Environmental Medicine<br />

Philadelphia, PA, USA<br />

Although many people assume that the bite of carbonation is due<br />

to tactile stimulation of the oral cavity by bubbles, it has become<br />

increasingly clear that carbonation bite comes mainly from<br />

<strong>for</strong>mation of carbonic acid in the oral mucosa. In Experiment<br />

1, we asked whether bubbles were in fact required to perceive<br />

carbonation bite. Subjects rated oral pungency from several<br />

concentrations of carbonated water both at normal atmospheric<br />

pressure (at which bubbles could <strong>for</strong>m) and at 2.0 atmospheres<br />

pressure (at which bubbles did not <strong>for</strong>m). Ratings of carbonation<br />

bite under the two pressure conditions were essentially identical,<br />

indicating that bubbles are not required <strong>for</strong> pungency. In<br />

Experiment 2, we created controlled streams of air bubbles<br />

around the tongue in mildly pungent CO 2<br />

solutions to determine<br />

how tactile stimulation from bubbles affects carbonation bite.<br />

Since innocuous sensations like light touch and cooling often<br />

suppress pain, we predicted that bubbles might reduce rated bite.<br />

Contrary to prediction, air bubbles flowing around the tongue<br />

significantly enhanced rated bite, without inducing perceived<br />

bite in blank (un-carbonated) solutions. Accordingly, though<br />

bubbles are clearly not required <strong>for</strong> carbonation bite, they may<br />

well modulate perceived bite. More generally, the results show<br />

that innocuous tactile stimulation can enhance chemogenic<br />

pain. Possible physiological mechanisms are discussed.<br />

Acknowledgements: Supported in part by Anheuser-Busch InBev<br />

#P125 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Acid detection by TRPV1 channels in both ‘taste blind’<br />

(P2X-KO) and C57 mice<br />

Meghan L Bills, Jennifer M Strat<strong>for</strong>d, Thomas E Finger<br />

University of Colorado School of Medicine/Department of Cell<br />

and Developmental Biology Aurora, CO, USA<br />

In the oropharynx the low pH of acidic solutions is detected<br />

via taste as sour, and by general mucosal fibers as pungency or<br />

chemesthesis. This results in the behavioral avoidance of acidic<br />

stimuli but the relative contribution of these two systems is<br />

unknown. Genetic deletion of the purinergic receptors P2X2<br />

and P2X3 (P2X-KO) results in interruption of taste bud-tonerve<br />

transmission and consequent loss of taste responses in<br />

the gustatory nerves. Although P2X-KO mice do not respond<br />

behaviorally to most tastants, they continue to avoid acids at<br />

similar concentrations as wildtypes. General mucosal afferents<br />

express TRPV1 channels, which are activated by low pH and<br />

may underlie this avoidance of acids. To test this, P2X-KO<br />

(n=8) and C57 (n=12) mice were assessed using a two-bottle<br />

preference test in which one bottle contained 20 mM citric acid<br />

(CA) and the other water. The test was given in the presence and<br />

absence of the TRPV1 antagonist Iodoresiniferatoxin (I-RTX).<br />

Both strains showed a significant decrease in avoidance with<br />

I-RTX versus vehicle (p


#P126 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

#P127 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

N-geranylcyclopropylcaboximide (NGCC) selectively activate<br />

hTRPV1 and hTRPA1 in cultured cells<br />

MR Rhyu 1 , HJ Son 1 , Y Kim 1 , MJ Kim 1 , SH Song 1 , MJ Cheong M 1 ,<br />

T Misaka 2 , M.L. Dewis 3 , V Lyall 4<br />

1<br />

Korea Food Research Institute Seongnam-Si, South Korea, 2 The<br />

University of Tokyo Tokyo, Japan, 3 International Flavors & Fragrances<br />

Union Beach, NJ, USA, 4 Virginia Commonwealth University<br />

Richmond, VA, USA<br />

In mammals, two salt taste pathways have been characterized:<br />

one is selectively responsive to Na + , which is inhibited by<br />

amiloride and the other is Na + non-specific and amilorideinsensitive.<br />

As the amiloride-sensitive Na + specific salt taste<br />

receptor, the epithelial sodium channel (ENaC) has been<br />

validated. The transient receptor potential vanilloid-1 variant salt<br />

taste receptor (TRPV1t) has been proposed as a constitutively<br />

active non-selective cation channel that has many similarities<br />

with the pain receptor TRPV1. In previous report we have<br />

shown that NGCC synthesized by IFF, modulates salt taste on<br />

human and amiloride-insensitive NaCl chorda tympani taste<br />

nerve responses by interacting with TRPV1t. In this presentation,<br />

we per<strong>for</strong>med calcium imaging and cell based assay using in<br />

hTRPV1-expressing cells to test the interaction with NGCC and<br />

TRPV1NGCC enhanced Ca 2+ influx in hTRPV1-exrpessing<br />

cells in a time- and dose-dependent manner with an EC 50<br />

value<br />

of 98.7 µM. The NGCC-induced Ca 2+ influx was markedly<br />

attenuated by ruthenium red (30 µM), a general blocker of<br />

TRP channels, and capsazepine (5 µM), a specific antagonist<br />

of TRPV1, implying NGCC directly activate TRPV1. On<br />

the other hand, TRPA1 is often co-expressed with TRPV1 in<br />

sensory neurons there<strong>for</strong>e we investigated the effects of NGCC<br />

on hTRPA1-expressing cells. NGCC enhanced Ca 2+ influx in<br />

hTRPA1-exrpessing cells in the same manner as in hTRPV1 with<br />

an EC50 value of 57.2 µM. The NGCC-induced Ca 2+ influx was<br />

blocked by ruthenium red (30 µM), and HC-030031 (100 µM), a<br />

specific antagonist of TRPA1. These data provides evidence that<br />

NGCC selectively activate TRPV1 and TRPA1 in cultured cells.<br />

These data further support our previous suggestion that NGCC<br />

interact with the TRPV1 variant cation channel in the anterior<br />

taste receptive field. Acknowledgements: Supported by a Korea<br />

Food Research Institute (KFRI) grant E0121201 and DC-011569<br />

Olfactory Overshadowing: The Effect of Verbal and Tactile<br />

Stimuli on Olfactory Memory<br />

Nicole K Beers, Amy E Callaham, David E Hornung<br />

Biology Dept. St. Lawrence University Canton, NY, USA<br />

While the strong association between smell and memory has<br />

received considerable attention, little is understood about the<br />

effect non-olfactory stimuli have on the encoding of olfactory<br />

memory. In the “verbal overshadowing” part of this study,<br />

subjects were divided into 5 groups which differed in the type of<br />

verbal response attached to 10 target odors presented during a<br />

training phase. In the 4 experimental groups, subjects verbalized<br />

names or descriptors whereas subjects in the control group<br />

provided no verbal name or description. After participating<br />

in an unrelated task, subjects were presented with a battery<br />

of odorants one at a time (10 targets and 10 distracters) and<br />

asked if they remembered smelling each odorant. Subjects who<br />

verbalized a name or description recognized fewer odorants<br />

as compared to subjects who provided no verbal response. In<br />

addition, the subjects who provided no verbal response identified<br />

fewer distracter odorants as belonging to the target group. In the<br />

“temperature overshadowing” part of this study, subjects had<br />

cold or room temperature water applied to their hands while they<br />

concurrently smelled the target odors. Subjects who had water<br />

applied to their hands during the training phase recognized fewer<br />

of the target odorants as compared to subjects in a control group<br />

from the verbal overshadowing part of the study who had no<br />

water applied to their hands; additionally subjects recalled fewer<br />

of the odors that were paired with the cold stimulus than those<br />

that were paired with room temperature water. The data from<br />

both parts of this study is consistent with the hypothesis that<br />

olfactory memory is impaired when target odorants are coupled<br />

with verbal or tactile stimuli. Acknowledgements:<br />

The St. Lawrence University Fellows Program and the Biology<br />

Department provided some of the funding <strong>for</strong> this work.<br />

#P128 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

A new paradigm <strong>for</strong> testing olfactory memory per<strong>for</strong>mance in<br />

healthy humans<br />

Yvonne F. Brünner 1 , Christian Benedict 2 , Jessica Freiherr 1<br />

1<br />

Diagnostic and Interventional Neuroradiology Aachen, Germany,<br />

2<br />

Institute of Neuroscience Uppsala, Sweden<br />

Long-term memory processes can be divided into two categories:<br />

declarative and procedural memory. While declarative memory<br />

concerns actual knowledge of facts, procedural memory applies<br />

to more unconscious memory of abilities and skills. Declarative<br />

memory processes are based upon the two components<br />

recollection and recognition. Odors are considered highly<br />

salient but abstract evocative cues during the recollection of<br />

past personal experiences and events. The aim of our current<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

77


study was to develop a paradigm <strong>for</strong> testing declarative memory<br />

processes with regard to odors. There<strong>for</strong>e, we implemented<br />

three memory tasks in E-prime 2.0 presentation software. With<br />

help of an odor-place associative memory task we investigated<br />

memory of odor-place combinations. During an odor item<br />

recognition task the subjects were asked if they had experienced<br />

several odors during the previous task. As a control task <strong>for</strong> the<br />

first task a picture-place associative memory task was utilized.<br />

Each of those tasks contained two phases – an encoding phase (1<br />

block) and a retrieval phase (4 blocks) – and were applied to 17<br />

healthy subjects. The results suggest that the subjects were able<br />

to learn during the encoding as well as during the retrieval phase.<br />

We found higher odor-place associative, odor recognition, and<br />

picture-place associative memory per<strong>for</strong>mance scores at the end<br />

compared to the beginning of the retrieval phase. We thus claim<br />

that our paradigm renders useful during the investigation of<br />

olfactory memory processes and the influence of external factors<br />

on those processes in humans. We here present a fully automated<br />

paradigm, which can be utilized during future functional imaging<br />

studies with the goal to shed light onto the neural network of<br />

human olfactory memory processes. Acknowledgements: This<br />

research was funded by a startup grant from the Medical Faculty<br />

of RWTH Aachen University.<br />

groups of at least 10 Ss on over a dozen such materials.<br />

We used the CPL protocols that have given stable results and<br />

required no normalization. The outcome fell in line with the<br />

data <strong>for</strong> the hundreds of materials, with neither more nor less<br />

variance. It adds evidence that the LSER has universal relevance<br />

to predict potency. Since the solvation-relevant parameters<br />

used <strong>for</strong> prediction exist <strong>for</strong> thousands of materials, the LSER<br />

may af<strong>for</strong>d more accessibility and accuracy than values in<br />

compilations. Acknowledgements: Supported by International<br />

Flavors and Fragrances.<br />

#P130 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Categorical Dimensions of Human Odor Descriptor Space<br />

Revealed by Non-Negative Matrix Factorization<br />

Jason B Castro 1 , Arvind Ramanathan 2 , Chakra S Chennubhotla 3<br />

1<br />

Bates College, Department of Psychology, Program in Neuroscience<br />

Lewiston, ME, USA, 2 Oak Ridge National Laboratory Oak Ridge, TN,<br />

USA, 3 University of Pittsburgh, Department of Computational and<br />

Systems Biology Pittsburgh, PA, USA<br />

#P129 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Fragrance Materials Extend the Range of a Model <strong>for</strong><br />

Odor Potency<br />

William S Cain 1 , Roland Schmidt 1 , Jorge E Cometto-Muñiz 1 ,<br />

Sang Park 1 , Craig B Warren 1 , Michael H Abraham 2 , Matthias Tabert 3<br />

1<br />

Chemosensory Perception Lab (CPL) La Jolla, CA, USA, 2 University<br />

College London / Chemistry London, United Kingdom, 3 International<br />

Flavors & Fragrances Union Beach, NJ, USA<br />

Olfactory research offers few trustworthy odor thresholds.<br />

Bad <strong>for</strong> the field, the situation sets nonexperts in need of the<br />

in<strong>for</strong>mation adrift. A rule of thumb says to choose the lowest<br />

threshold of a set because studies with the best stimulus control<br />

and some measure of concentration normally yield the lowest<br />

values. Nagata et al. over years produced the largest set of<br />

thresholds gathered with coherent methodology and analytical<br />

verification. They lie below most others. Abraham et al. used a<br />

well-established linear solvation energy relationship (LSER) to<br />

describe the set. To strengthen the position, they incorporated<br />

sets from the CPL and Hellman. Data obtained via olfactometer<br />

in recent years from the CPL needed no normalization with the<br />

Nagata set, whereas those from Hellman and earlier squeezebottle<br />

data from the CPL did. The LSER accommodated these<br />

via indicator variables (factors) to bring the data into line and<br />

described potency <strong>for</strong> 353 odorants. Although the equation left<br />

about 25% of variance unaccounted <strong>for</strong>, it offers the strongest<br />

statement yet <strong>for</strong> prediction. Until now, no sets included<br />

fragrance materials, such as patchouli oil or vanillin, known <strong>for</strong><br />

potency. Without them, one could argue that the equation might<br />

describe local rather than universal rules. We accordingly ran<br />

Recent studies using Principal Components Analysis (PCA)<br />

support low-dimensional models of odor space, in which one or<br />

two dimensions – with hedonic valence featuring prominently –<br />

explain most odor variability. Here we use non-negative matrix<br />

factorization (NMF) – a nonlinear optimization method - to<br />

discover an alternative, reduced-dimensional representation<br />

of the Dravnieks odor database(144 odors x 146 descriptors).<br />

We first divided the dataset into training and testing halves,<br />

and found that RMSD testing error attained a minimum <strong>for</strong><br />

subspace choice of 25, motivating this as an upper bound <strong>for</strong><br />

odor perceptual space dimensionality. More parsimonious<br />

representations were found by comparing reconstruction errors<br />

(fraction of unexplained variance) of NMF with reconstruction<br />

errors of PCA on scrambled data (PCAsd). For subspace<br />

sizes > 10, NMF error was indistinguishable from PCAsd<br />

error, indicating no gain in retaining more than 10 perceptual<br />

dimensions. As is typical of NMF basis sets, the 10 odor<br />

dimensions we obtain are sparse (only a small subset of the<br />

146 descriptors apply), and categorical (represent a positive<br />

valued quality). Moreover, these 10 dimensions were nearorthogonal,<br />

with a mean angle of 73 degrees between all pairs<br />

of basis vectors. Investigating the distribution of odors in this<br />

10-dimensional space, we find marked clustering, with each<br />

odor clearly defined by its membership in a single dimension, to<br />

the exclusion of others. Members of each cluster have notable<br />

structural homology, which we quantified as correlations among<br />

physiochemical descriptors of odors in each cluster. In sum, we<br />

describe a representation of odor perceptual space consisting of<br />

10 discretely occupied dimensions that apply categorically. We<br />

propose that this may help elucidate the natural axes of olfaction.<br />

Acknowledgements: NIH GM086238 (CSC)<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

78


#P131 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Genetic Contribution to Binaral Rivalry<br />

Kepu Chen, Bin Zhou, Guo Feng, Wen Zhou<br />

Institute of Psychology, Chinese Academy of <strong>Sciences</strong> Beijing, China<br />

When two odorants of different structure and smell are<br />

simultaneously presented to the two nostrils, we experience<br />

alternations in olfactory percepts, a recently discovered<br />

phenomenon termed binaral rivalry. In an ef<strong>for</strong>t to further<br />

characterize this phenomenon and its nature, we adopted the<br />

twin method and tested monozygotic (MZ) twins (n = 73 pairs)<br />

that are genetically identical and dizygotic (DZ) twins (n =<br />

70 pairs) that share about half of their genes. The majority of<br />

participants experienced binaral rivalry over a course of 20<br />

samplings of eugenol and amyl acetate, one to each nostril.<br />

Large variances are observed in the number and the magnitude<br />

of their perceptual switches. Critically, such individual<br />

differences are partially genetic. The correlations between MZ<br />

twins <strong>for</strong> the two a<strong>for</strong>ementioned indexes are both higher than<br />

those between DZ twins. The best-fitting genetic models showed<br />

that over 30% of the variances in binaral rivalry rate and binaral<br />

rivalry magnitude, respectively, were accounted <strong>for</strong> by additive<br />

genetic factors. Our study represents the first large sample study<br />

of binaral rivalry. The findings demonstrate a reliable genetic<br />

component of this phenomenon and suggest an innate rhythm of<br />

olfactory perception.<br />

#P132 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Olfactory Plasticity in Young Adults<br />

Beverly J. Cowart 1 , Marcia L. Pelchat 1 , Johan N. Lundström 1,2,3 ,<br />

Ryan Craw<strong>for</strong>d 1 , Lydia Milbury 1 , Kathrin Ohla 1,4<br />

1<br />

Monell Chemical Senses Center Philadelphia, PA, USA, 2 Dept. of<br />

Psychology, University of Pennsylvania Philadelphia, PA, USA,<br />

3<br />

Dept. of Clinical Neuroscience, Karolinska Institute Stockholm,<br />

Sweden, 4 German Institute of Human Nutrition Potsdam-Rehbrücke<br />

Nuthetal, Germany<br />

There is considerable evidence that olfaction is, in many ways,<br />

a “learned” sense, showing experience-induced plasticity in<br />

both central circuits and peripheral receptors even in adulthood.<br />

Although this has been demonstrated in humans, the focus<br />

has been on single chemicals to which some individuals are<br />

initially insensitive. We sought to explore (1) the possibility of<br />

enhancing young adult sensitivity to complex odors, (2) the role<br />

of cognitive engagement (active identification of exposed odors<br />

vs. simple exposure), and (3) the extent of transfer of learning to<br />

unexposed odors. For this, we obtained both odor thresholds and<br />

olfactory event-related potentials (OERPs) from 40 young adults<br />

in an attempt to assess the neuronal mechanisms of potential<br />

behavioral changes. Thresholds <strong>for</strong> four exposed and two<br />

unexposed complex odorants were obtained at baseline, 6 and<br />

12 weeks; OERPs in response to two of the exposed and both<br />

unexposed odorants were obtained at baseline and 12 weeks.<br />

Our behavioral results suggest that intermittent odor exposure<br />

in these circumstances does enhance threshold sensitivity and<br />

that cognitive engagement may further enhance generalization<br />

to unexposed odorants. Our electrophysiological results show<br />

learning-dependent amplitude changes, particularly of the late<br />

positive component. Taken together, these data provide further<br />

support <strong>for</strong> the notion that repeated exposure augments olfactory<br />

sensitivity and that cognitive mechanisms exert a significant<br />

modulatory effect. Acknowledgements: Supported by the U.S.<br />

Army Research Office, grant #W911NF-11-1-0087.<br />

#P133 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Proton-Transfer-Reaction Mass Spectrometry<br />

Melanie Y Denzer 1 , Jonathan Beauchamp 2 , David W Kern 3 , Stefan<br />

Gailer 2 , Norbert Thuerauf 4 , Johannes Kornhuber 4 , Andrea Buettner 1,2<br />

1<br />

Department of Chemistry and Pharmacy, Emil Fischer Center,<br />

University of Erlangen-Nuremberg Erlangen, Germany, 2 Department<br />

of Sensory Analytics, Fraunhofer Institute <strong>for</strong> Process Engineering<br />

and Packaging IVV Freising, Germany, 3 Department of Comparative<br />

Human Development, Institute <strong>for</strong> Mind and Biology, University<br />

of Chicago Chicago, IL, USA, 4 Department of Psychiatry and<br />

Psychotherapy, University of Erlangen Erlangen, Germany<br />

Since their introduction in the mid-1990s, Sniffin’ Sticks have<br />

been used effectively by many otolaryngologists to assess<br />

olfactory dysfunction in countless patients. Despite their<br />

widespread use, however, there is currently a lack of data on<br />

the actual odorant concentrations released from the tips of<br />

these pens and whether these emitted concentrations scale<br />

linearly in accordance with the odorant concentrations of the<br />

pen set. The purpose of this study was to ascertain whether the<br />

Sniffin’ Sticks’ presumed odorant release was concordant with<br />

the concentration of the odorant solutions placed in the pens.<br />

The commercially-available odour threshold test containing<br />

n-butanol was chosen here <strong>for</strong> evaluation. The threshold set<br />

contains concentrations (v/v) ranging from 4 % (pen no. 1)<br />

to 1.2 ppm v<br />

(pen no. 16), with stepwise 1:2 dilutions. We also<br />

tested an additional custom-made pen containing 8 % n-butanol<br />

(pen no. 0). The odorant concentration emanating from the tip<br />

of each pen was measured directly via proton-transfer-reaction<br />

mass spectrometry (PTR-MS), which is an on-line analytical tool<br />

<strong>for</strong> detection and quantitation of volatile organic compounds<br />

(VOCs) – including odorants – at trace concentrations. The pens<br />

were also subjected to repeated use to ascertain the degree of<br />

reproducibility of emitted odorant concentrations under stress.<br />

These measurements showed that the concentration linearity of<br />

n-butanol emitted over the range of pens was excellent and highly<br />

reproducible. The stress tests demonstrated that the emitted<br />

concentrations of n-butanol were lower after repeated use of the<br />

pens compared to those of the unused pens, albeit with a mostly<br />

good linearity over the entire range of pens. Acknowledgements:<br />

Part of this study is affiliated with the Neurotrition Project,<br />

which is supported by the FAU Emerging Fields Initiative. This<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

79


work was also partly supported by The National Social Life,<br />

Health and Aging Project Wave 2 (R37 AG030481). DWK is<br />

supported by The Center on Aging Specialized Training Program<br />

in the Demography and Economics of Aging, which is funded by<br />

the National Institute on Aging (NIA) (T32000243)<br />

#P134 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Effects of Cinnamon Scent Administration on Physiology,<br />

Range of Motion, Mood, Anxiety and Perceived Workload<br />

During a Multi-session Physical Therapy Program<br />

Jessica Florian, Kristen Johnson, Sierra Moore, Bryan Raudenbush,<br />

Allison Burke<br />

Wheeling Jesuit University Wheeling, WV, USA<br />

Scents have been shown to elicit both emotional and<br />

physiological responses. The current study aimed to evaluate the<br />

possible effects of cinnamon scent when applied to a physical<br />

therapy regimen. Forty-two undergraduate students, 16 males<br />

and 26 female, completed a four trial physical therapy regimen<br />

in one of two rooms: a control room or a room infused with<br />

a cinnamon scent. The experimenters measured participants’<br />

range of motion, mood (POMS), and anxiety (STAI) prior to<br />

and following each trial of exercises. At the end of each visit,<br />

perceived workload was assessed (NASA-TLX). The data were<br />

analyzed using a 4 (visits) X 2 (groups) mixed design ANOVA.<br />

Significant results were found <strong>for</strong> ratings of ef<strong>for</strong>t on the NASA-<br />

TLX, F(3, 120)=2.8, p = .042. Participants in the cinnamon scent<br />

condition rated their perceived ef<strong>for</strong>t exertion as being lower<br />

than participants in the control condition. Decreased perceived<br />

ef<strong>for</strong>t may cause patients undergoing a physical therapy program<br />

to feel more com<strong>for</strong>table while completing their exercises, thus<br />

increasing the likelihood of adherence to the program.<br />

#P135 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Towards A Novel Method <strong>for</strong> Human Olfactory Research<br />

Jessica M Gaby, Vivian Zayas<br />

Cornell University Ithaca, NY, USA<br />

The majority of human olfactory research to date has been<br />

conducted without regard to perfume or dietary contributions to<br />

body odor. Given that humans have used perfume <strong>for</strong> thousands<br />

of years, and that culturally mediated food preferences may<br />

affect body odor, I propose that an ecologically relevant model<br />

of olfactory communication should account <strong>for</strong> body odor as<br />

people present themselves in real social situations. Further,<br />

current collection methods <strong>for</strong> odor samples focus mainly on<br />

axillary secretions, though it is known that other body areas<br />

contribute to odor. This pilot study aims to validate a novel<br />

olfactory research method by employing present others rather<br />

than disembodied odors. In a repeated measures design,<br />

blindfolded, ear-plugged raters evaluated a series of participants<br />

as each sat beside them. No restrictions in diet or fragrance were<br />

used, to mimic real-life interactions. First and second ratings <strong>for</strong><br />

each participant were examined <strong>for</strong> intertrial reliability, revealing<br />

significant consistency (p


cortex during control versus prediction is currently underway and<br />

will also be discussed. Acknowledgements: The research in this<br />

presentation is supported by the National Institutes of Health<br />

[NIDCD 1 R01 DC 010014, NIMH 5 F32 MH 091967].<br />

#P137 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Pleasant and unpleasant odors are easier to detect and<br />

harder to ignore<br />

Jingwen Jin, Aprajita Mohanty<br />

State University of New York at Stony Brook/Department of<br />

Psychology Stony Brook, NY, USA<br />

Odor objects are often experienced within the context of several<br />

competing odors. To deal with the excess of in<strong>for</strong>mation the<br />

olfactory system utilizes mechanisms that bias the competition<br />

between odors towards preferential representation of the most<br />

relevant odor. This biasing may involve anticipatory attentional<br />

mechanisms that use feature-related in<strong>for</strong>mation regarding<br />

behaviorally relevant odors to amplify relevant odors and filter<br />

our irrelevant odors. Since an important feature of olfactory<br />

stimuli is their valence, the present study utilized an olfactory<br />

attentional search task to examine whether 1) expectation of<br />

pleasant or unpleasant odors enhances perceptual sensitivity<br />

compared to neutral odors and 2) presence of an irrelevant<br />

pleasant or unpleasant odor make it harder to detect a relevant<br />

odor. Subjects per<strong>for</strong>med an olfactory task in which they decided<br />

whether a particular target smell is present in each trial. Trials<br />

started with a cue that indicated the <strong>for</strong>thcoming target odor<br />

(pleasant odor A, unpleasant odor B, or neutral odor C) followed<br />

by the target stimulus. The stimulus consisted of either odor A<br />

alone, odor B alone, odor C alone, or binary mixture of these<br />

odors (AB, AC, and BC). In our results,d-prime, a measure of<br />

perceptual sensitivity showed the following trend: pleasant odor<br />

following pleasant cues > unpleasant odors following unpleasant<br />

cues > neutral odors following neutral cues (F(1,13)=7.240,<br />

p


3.3 mm) placed at the olfactory cleft under endoscopic guidance.<br />

Participants did not require topical anaesthetics or decongestants<br />

and sniffed naturally (n=35; 21 female, 32±7 years of age).<br />

An additional cannula fitted at the nostrils was connected to<br />

a pressure transducer to record inspiratory nasal pressure and<br />

breathing cycles throughout odorant presentation. Following a<br />

protocol similar to that which is used during clinical assessments,<br />

n-butanol was presented in a triad of Sniffin’ Sticks pens, 1<br />

odorant target pen and 2 ‘odorless’ blanks. Participants were<br />

asked to identify the odorant-containing target pen in each of six<br />

triads with increasing n-butanol concentrations (0.125, 0.25, 0.5,<br />

2, 4, and 8 %). The entire protocol only required approximately<br />

15 minutes to per<strong>for</strong>m. All participants were fully compliant and<br />

reported only minimal discom<strong>for</strong>t. The catheter and cannula did<br />

not disrupt normal breathing patterns. This protocol allowed us<br />

to quickly and effectively measure odorant concentrations at the<br />

olfactory cleft in real time. Comparing the known concentration<br />

of the stimulus odorant and the concentration found at the<br />

olfactory cleft will elucidate how to best adjust and develop<br />

measures designed to determine an individual’s threshold<br />

sensitivity. Acknowledgements: This work was supported in part<br />

by The National Social Life, Health and Aging Project Wave 2<br />

(R37 AG030481), and is affiliated with the Neurotrition Project,<br />

supported by the FAU Emerging Fields Initiative. DWK is<br />

supported by The Center on Aging Specialized Training Program<br />

in the Demography and Economics of Aging (National Institute<br />

on Aging (T32000243). TH received funding from the DFG<br />

(HU 441/10-1).<br />

#P140 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Quantitative classification of the perceived intensity curve<br />

of a continuously presented odor<br />

Tatsu Kobayakawa 1 , Tomoko Matsubasa 2 , Naomi Gotow 1<br />

1<br />

The National Institute of Advanced Industrial Science and Technology<br />

(AIST) Ibaraki Tsukuba, Japan, 2 Tokyo Gas Co., Ltd Tokyo, Japan<br />

Researchers of olfaction have devoted significant ef<strong>for</strong>t to<br />

understanding the phenomenon of adaptation. Previous<br />

studies of olfactory adaptation have reported, in general, that<br />

perceived intensity of a continuously presented odor decreases<br />

exponentially over time. More recently, however, other studies<br />

have shown that perceived intensity of continuous odor does<br />

not always decrease exponentially. Some studies focused on the<br />

various time-course of perceived intensity of a continuously<br />

presented odor, whereas few studies have tried to classify<br />

time-course patterns quantitatively. In this study, we per<strong>for</strong>med<br />

real-time evaluation of perceived intensity of a continuously<br />

presented odor <strong>for</strong> 480 seconds using nine odorants, and<br />

quantitatively classified the perceived intensity curves using the<br />

time points <strong>for</strong> 20-80% of time integral of perceived intensity.<br />

This analysis revealed that intensity curves could be classified<br />

into two clusters: one group (“typical”) exhibited the commonly<br />

observed exponential decrease in perceived intensity; the other<br />

group (“non-typical”) included a range of time courses rather<br />

than strict exponential decrease. The average intensity of<br />

“typical” group decreased below threshold three minutes after<br />

beginning of odor exposure, and that of “non-typical” group,<br />

on the other hand, did not decrease under threshold during total<br />

measuring time.<br />

#P141 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Effects of Peppermint Scent Administration on Augmenting<br />

Cognitive and Creative Per<strong>for</strong>mance<br />

Lucas Lemasters, August Capiola, Bryan Raudenbush, Sierra Moore<br />

Wheeling Jesuit University Wheeling, WV, USA<br />

Level of creativity has been assessed in a number of ways,<br />

including interactions between body and environment on creative<br />

thinking. Environmental richness has been shown to interact<br />

with creativity, with greater levels of environmental richness<br />

leading to more creative responses. The present study attempted<br />

to determine if peppermint scent administration could promote<br />

creativity, since past research with peppermint scent reports<br />

improved per<strong>for</strong>mance on clerical tests, thus hinting at the<br />

possibility <strong>for</strong> cognitive enhancement. Participants completed<br />

the Torrance® Tests of Creative Thinking, a standardized test<br />

measuring creative thinking abilities, in both a non-scented<br />

condition (control) and a peppermint scented condition.<br />

Different versions of the test, as well as the conditions, were<br />

counterbalanced. The data were subjected to paired samples<br />

t-tests, with condition (peppermint, control) serving as the<br />

independent measure and raw scores of fluency, originality,<br />

elaboration, abstractness-of-titles, and resistance-to-prematureclosure<br />

serving as dependant measures. There was a significant<br />

difference between the conditions <strong>for</strong> fluency [t(33)=-2.41,<br />

p=.02], originality [t(33)=-2.13, p=.04], and elaboration<br />

[t(33)=-7.38, p=.00], with all measures having higher scores<br />

<strong>for</strong> the peppermint scent condition. Implications suggest<br />

working conditions <strong>for</strong> those individuals with occupations that<br />

require creative thinking and problem solving may benefit from<br />

peppermint scented working conditions.<br />

#P142 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Contextual effects on hedonic evaluation of odors<br />

Shiori Nakano, Saho Ayabe-Kanamura<br />

Univ. of Tsukuba Tsukuba, Japan<br />

When we sequentially evaluate a character of sensory modality<br />

stimuli, evaluation of the stimuli might be influenced by what<br />

precedes them; contextual effects. Hedonic contrast, is one of<br />

these phenomenon, when a stimulus is preceded by a more<br />

pleasant stimulus, its pleasantness is rated lower (negative<br />

contrast), whereas when a stimulus is preceded by a less pleasant<br />

stimulus, its pleasantness is rated higher (positive contrast).<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

82


In this study, we investigated the contrast effect during evaluation<br />

of odor intensity and pleasantness. Two sets of positive odors<br />

and a set of negative odors were used. Group 1 sequentially rated<br />

the positive set A, the negative set, and the positive set B. Group<br />

2 rated the negative set and then the positive set B. Group 3 rated<br />

only the positive set B. The mean rating of the negative and the<br />

positive set B were each compared among groups. As a result, in<br />

the intensity rating, both positive and negative contrast occurred<br />

clearly. In the pleasantness rating, only negative contrast was<br />

seen. This suggested that, in olfaction, the influence of the<br />

positive context stimuli on the following negative stimuli might<br />

be more robust. To investigate the cause of the non-occurrence<br />

of positive contrast, we analyzed a change of hedonic evaluation<br />

by rating order within a positive or negative set. While there was<br />

no influence of the rating order in a negative set, the pleasantness<br />

rating of a positive set tended to gradually decrease. This result<br />

is discussed in terms of the unstable characteristics <strong>for</strong> hedonic<br />

evaluation of positive odors and the stability of the hedonic<br />

rating of negative odors. There<strong>for</strong>e, positive hedonic contrast<br />

may not occur due to the hedonic value of negative odors being<br />

carried over to the following rating of the better stimuli.<br />

was observed while the subjects were asleep. Although lavender<br />

has a sedative effect on humans thorough the activation of<br />

PSNS, inhaling this popular aroma during sleep may have a<br />

significant impact on cortisol secretion by enhancing HPA<br />

activity after awakening.<br />

#P144 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Common aversive learning mechanisms to olfactory and<br />

visual stimuli<br />

Valentina Parma 1,2 , Stacie S. Miller 1,3 , Fredrik Åhs 4 ,<br />

Johan N. Lundström 1,5,6<br />

1<br />

Monell Chemical Senses Center Philadelphia, PA, USA,<br />

2<br />

Department of General Psychology, University of Padova Padova,<br />

Italy, 3 International Flavor and Fragrances Union Beach, NJ, USA,<br />

4<br />

Center <strong>for</strong> Cognitive Neuroscience, Duke University Durham, NC,<br />

USA, 5 Department of Clinical Neuroscience, Karolinska Instituet<br />

Stockholm, Sweden, 6 Department of Psychology, University of<br />

Pennsylvania Philadelphia, PA, USA<br />

#P143 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Increased cortisol secretion after awakening following<br />

lavender inhalation during sleep<br />

Shusaku Nomura 1 , Masako H. Ohira 2 , Toyonari Fujikawa 1 ,<br />

Kanetoshi Ito 3<br />

1<br />

Nagaoka University of Technology/Engineering Nagaoka,<br />

Japan, 2 Shiga University/Education Otsu, Japan, 3 TAKASAGO<br />

INTERNATIONAL CORPORATION Kanagawa, Japan<br />

Aromatherapy improves sleep through autonomous nervous<br />

system (ANS) modulation. However, few studies have focused<br />

on the effect of olfactory exposure during sleep on hormonal<br />

secretion. The objective of this study was to investigate the<br />

effect of inhaling odorants on cortisol secretion during sleep<br />

and after awakening. Salivary cortisol was assessed because<br />

this glucocorticoid represents activation of the hypothalamuspituitary-adrenal<br />

(HPA) system, which is a dominant stress<br />

reaction pathway in the body. We used essential oils of jasmine<br />

and lavender, which induce activation of the sympathetic and<br />

parasympathetic nervous systems (PSNS), respectively. These<br />

oils were administered to subjects through an olfactometer;<br />

volatilized odorants were intermittently delivered through a<br />

cannula (first 1 min of each 5 min interval) during a 6-h sleep<br />

period. Subjects experienced each odorant condition [jasmine,<br />

lavender, or scentless air (control) each night] and were exposed<br />

to all three conditions in a counter-balanced order. Saliva<br />

samples were collected every 30 min while the subjects were<br />

asleep using our own-developed saliva collection equipment<br />

and every 15 min <strong>for</strong> 1 h after awakening. Surprisingly, salivary<br />

cortisol after awakening was significantly higher under the<br />

lavender condition than in the jasmine (p


#P145 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Perception of large and small odorant molecules –<br />

A study investigating the olfactory perception as a<br />

function of age and odorant molecular size<br />

Laura Puschmann, Charlotte Sinding, Thomas Hummel<br />

Smell & Taste Clinic, Department of Otorhinolaryngology,<br />

University of Dresden Medical School Dresden, Germany<br />

For various sensory systems selective functional limitations with<br />

increasing age were detected. In vision and hearing it shows in<br />

hyperopia or an increased hearing threshold <strong>for</strong> high frequencies.<br />

But what about the olfactory system? While some studies suggest<br />

a general olfactory impairment, others allow assumptions of<br />

partial processes. In this study the influence of odorant molecular<br />

size was examined <strong>for</strong> the olfactory perception in different ages.<br />

Olfactory threshold tests were conducted at two age groups<br />

(group 1: 18 to 30 years, group 2: 50 to 70 years). We used single<br />

odorants, bimolecular odorants and perfumes, each consisting of<br />

large or small molecules. The results of these studies showed no<br />

differences <strong>for</strong> large and small odorant molecules in the young<br />

group of subjects. Whereas they were perceived differently by the<br />

50 to 70 year-old. The latter had significantly higher thresholds<br />

<strong>for</strong> large olfactory molecules. This phenomenon however, was<br />

detected only <strong>for</strong> mono- and bimolecular odorants. Perfumes<br />

were perceived similar in both groups. In Conclusion we can<br />

hypothesize differing processes <strong>for</strong> odor perception of small and<br />

big odorant molecules in different age-groups at the receptor<br />

level. Moreover perfumes seem to generate more complex<br />

olfactory in<strong>for</strong>mation. The existence of partial processes on<br />

olfactory impairment with increasing age is expected <strong>for</strong> the<br />

olfactory system, too. Acknowledgements: This project was<br />

supported by a grant from DFG Schwerpunktprogramm (SPP)<br />

1392 - Integrative Analysis of Olfaction to Thomas Hummel. In<br />

addition we thank Fragrance Resources GmbH, Hamburg <strong>for</strong><br />

cooperation.<br />

#P146 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Odor Identification and Cognition in a Nationally<br />

Representative Sample of Older Adults<br />

L. Philip Schumm 1 , David W. Kern 2 , Kristen E. Wroblewski 1 , Jayant<br />

M. Pinto 3 , Ashwin A. Kotwal 4 , William Dale 4 , Martha K. McClintock 2<br />

1<br />

Department of Health Studies, University of Chicago Chicago, IL,<br />

USA, 2 Comparative Human Development and Institute <strong>for</strong> Mind<br />

and Biology, University of Chicago Chicago, IL, USA, 3 Section of<br />

Otolaryngology-Head and Neck Surgery, Department of Surgery,<br />

University of Chicago Chicago, IL, USA, 4 Department of Medicine,<br />

Section of Geriatrics & Palliative Medicine, University of Chicago<br />

Chicago, IL, USA<br />

Errors in odor identification are associated with subsequent<br />

cognitive impairment in populations clinically at risk <strong>for</strong><br />

Alzheimer and Parkinson diseases. However, such associations<br />

have not been investigated in the general U.S. population.<br />

We administered a 5-item test of odor identification to a U.S.<br />

national probability sample of 3,005 community-dwelling adults<br />

aged 57–85 in 2005–6 (Wave 1); respondents and their spouses<br />

were then retested in 2010–1 (Wave 2). Odors were presented<br />

using Sniffin’ Sticks, and respondents were asked to identify<br />

each from among four word-picture options. Respondents<br />

in Wave 2 also completed the Chicago Cognitive Function<br />

Measure (C-CFM), a survey instrument assessing eight distinct<br />

cognitive domains, derived from the Montreal Cognitive<br />

Assessment (MoCA). The C-CFM is 30% faster, with scores<br />

ranging from 0–20 (mean = 13.5, SD = 4.1) that correlate highly<br />

with the MoCA (r = 0.97). In a multiple regression model of<br />

C-CFM on odor identification score in Wave 2 (n = 2,076),<br />

each additional odor identification error was associated with<br />

a reduction in C-CFM score of 0.61 (95% CI = (0.44, 0.77); p<br />


(Wave 2), assessing demographics, social life, and health,<br />

including olfaction (N = 1436). Odor identification was<br />

measured with 5 Sniffin’ Sticks (0-5 correct). Multivariate linear<br />

regression quantified the association between olfactory decline<br />

over 5 years and demographic, health and psychosocial factors.<br />

Odor identification declined most rapidly in older individuals<br />

(0.25 greater decline in number correct per decade of age,<br />

P0.05). In addition to<br />

having poor olfactory function, men and older people experience<br />

accelerated olfactory decline, effects not explained by our<br />

measured psychosocial or health conditions. We find no evidence<br />

of accelerated olfactory aging explaining the health disparity<br />

in Blacks seen at the Wave 1 and 2 time points, suggesting that<br />

major insults to the olfactory system occur be<strong>for</strong>e middle age.<br />

Acknowledgements: The National Institute on Aging (R37<br />

AG030481 AG036762 AG029795), the Institute <strong>for</strong> Translational<br />

Medicine at The University of Chicago (KL2RR025000), the<br />

McHugh Otolaryngology Research Fund, and the American<br />

Geriatrics Society.<br />

#P148 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Receptor Representations of Perceptual Similarity<br />

Andrew H. Moberly 1,2 , Lindsey L. Snyder 2 , Joel D. Mainland 1,2<br />

1<br />

University of Pennsylvania Philadelphia, PA, USA, 2 Monell Chemical<br />

Senses Center Philadelphia, PA, USA<br />

In the current consensus model of the olfactory code, the<br />

recognition of an odorant molecule depends on which receptors<br />

are activated and to what extent. Here we set out to determine<br />

if the receptor-activation profile <strong>for</strong> a set of camphoraceous<br />

odorants is more representative of the structural features or the<br />

perceptual similarity. The camphoraceous odor class is unusual<br />

in that despite their common perceptual odor character the<br />

odorants do not share a common functional group (median<br />

Tanimoto similarity = 0.14). Using a set of odorants described<br />

as having a camphoraceous quality (Amoore, 1970), we tested<br />

the similarity of the odorants in terms of molecular structure,<br />

receptor activation profile in a heterologous assay, and human<br />

perceptual ratings. Molecular structure was quantified using<br />

physicochemical descriptors (Haddad et al., 2008). To measure<br />

receptor-activation profiles, we cloned receptors representing<br />

384 of the most frequent variants in the 1000 genomes data<br />

set. We then tested the odorants against these receptors using a<br />

heterologous luciferase assay. Human perceptual similarity was<br />

assayed using an air-dilution olfactometer to obtain pairwise<br />

similarity ratings. Preliminary results suggest that the receptor<br />

array is more representative of structural features than of<br />

perceptual similarity. Acknowledgements: R03-DC011373<br />

#P149 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Taste-evoked Chorda Tympani Responses in C57BL/6J Mice<br />

Vary Depending on Which Region of the Tongue is Stimulated<br />

Rachel M. Dana, Stuart A. McCaughey<br />

Ball State University Muncie, IN, USA<br />

It is known that each taste bud is responsive to stimuli<br />

representing all of the basic taste qualities. However, there is<br />

also evidence that the transduction mechanisms that mediate<br />

such responses vary depending on location in the oral cavity,<br />

and the gustatory nerves differ in amiloride-sensitivity and other<br />

properties. Although the peripheral taste-responsive nerves<br />

have often been compared with each other, there have been few<br />

attempts to consider whether the properties of a particular nerve<br />

vary depending on which oral regions are stimulated with taste<br />

solutions. We there<strong>for</strong>e measured taste-evoked chorda tympani<br />

(CT) responses in mice while flowing solutions over the anterior<br />

or posterior tongue separately. Subjects were C57BL/6J mice,<br />

which were anesthetized prior to surgery to expose the CT. A<br />

silicone rubber ring was used to divide the tongue into anterior<br />

and posterior halves, and tastants were flowed selectively over<br />

one half at a time while measuring whole-nerve CT activity.<br />

Responses to NaCl and sucrose were significantly larger, but<br />

responses to HCl and quinine were smaller, when stimulating<br />

the anterior tongue relative to the posterior. The responses to<br />

NaCl mixed with amiloride, however, were similar <strong>for</strong> the two<br />

regions due to a greater suppressive effect of amiloride on the<br />

anterior as compared to posterior tongue. Our data suggest that<br />

the properties of CT fibers vary depending on the location of<br />

the taste buds that they innervate. The characteristics of our<br />

posterior tongue responses, in fact, were more similar to those<br />

normally associated with the glossopharyngeal nerve than with<br />

the CT. Additional work will be needed to delineate the relative<br />

contributions of fungi<strong>for</strong>m and foliate papillae to our posterior<br />

tongue responses.<br />

#P150 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Dried-bonito dashi: Taste qualities evaluated using CTA<br />

methods in wild type and T1R1 KO mice.<br />

Eugene R. Delay 1 , Takashi Kondoh 2<br />

1<br />

University of Vermont/Department of Biology Burlington, VT,<br />

USA, 2 Ajinomoto/Human Health and Nutrition Research Group<br />

Kawasaki, Japan<br />

Dried-bonito dashi, a broth used to increase the palatability of<br />

Japanese cuisine, is made from dried kelp, fish oils, shiitake<br />

mushrooms, and other sources. Flavor enhancement is due to<br />

olfactory and taste stimuli such as lactic acid, L-amino acids<br />

(including glutamate), and inositol monophosphate (IMP).<br />

Kawasaki et al. (2008) and Kondoh et al. (2012) report that rats<br />

and mice prefer dashi in 2-bottle preference tests. We conducted<br />

conditioned taste aversion (CTA) experiments to determine what<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

85


taste qualities mice can identify in dashi and if an L-amino acid<br />

receptor (T1R1) contributes to these sensations. C57BL/6J wild<br />

type (WT) mice and T1R1 knockout (KO) mice made against<br />

a mixed background and backcrossed to C57BL mice (Zhao et<br />

al., 2003) were used to see if a CTA to dashi generalized to (1)<br />

5 basic tastes, and (2) individual L-amino acids (+IMP, pH 5.7-<br />

5.8) found in dashi (Ajinomoto, Japan), with or without lactic<br />

acid added. The role of odor cues was minimized by constant<br />

exposure to the odor of dashi throughout the experiment and<br />

by compromising the nasal epithelium with ZnSO4 (verified<br />

by a “chip” odor test). Generalization of the CTA was greatest<br />

to sucrose and weakest to NaCl in WT mice. WT mice also<br />

generalized this CTA to individual amino acids to a degree<br />

roughly related to the concentration of the amino acid in dashi.<br />

Lactic acid (1%) altered generalization of the CTA to a subset of<br />

amino acids. KO mice showed a similar pattern of generalization<br />

to basic tastes but none to L-glutamate. Even though KO<br />

mice readily learned a CTA <strong>for</strong> dashi, they showed minimum<br />

generalization to the other L-amino acids. These results show<br />

that dashi elicits a complex taste and that the T1R1 receptor<br />

contributes significantly but not entirely to that complexity.<br />

Acknowledgements: NSF IOS-0951016<br />

#P151 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

GABA as afferent taste neurotransmitter<br />

Gennady Dvoryanchikov 1 , Elizabeth Pereira 1 , Christopher Williams 2 ,<br />

Stephen D. Roper 1,3 , Nirupa Chaudhari 1,3<br />

1<br />

Department of Physiology and Biophysics, University of Miami<br />

Miller School of Medicine Miami, FL, USA, 2 Graduate Program in<br />

Biomedical <strong>Sciences</strong>, University of Miami Miller School of Medicine<br />

Miami, FL, USA, 3 Program in Neurosciences, University of Miami<br />

Miller School of Medicine Miami, FL, USA<br />

ATP is recognized as a principal taste transmitter, activating<br />

P2X2 and P2X3 purinoceptors located on afferent sensory<br />

fibers that innervate taste buds. Taste buds also contain and<br />

release several other neurotransmitters. Many of these have<br />

been shown to mediate cell-to-cell communication within the<br />

bud. For instance, we recently reported that GABA serves as<br />

an inhibitory transmitter, decreasing taste-evoked ATP release<br />

from taste buds. We now ask whether GABA might also act<br />

on afferent sensory axons that innervate taste buds. To answer<br />

this question, we analyzed GABA receptors in geniculate<br />

ganglia. Neurons in this ganglion innervate fungi<strong>for</strong>m and<br />

palatal taste buds. Using RT-PCR, we show that GABA-A a 1<br />

is prominently expressed along with lower levels of a2, b2, b3,<br />

g1 and g2 subunits. The cation-chloride cotransporters, KCC3<br />

and KCC4, and particularly, NKCC1, which is essential <strong>for</strong> the<br />

inhibitory action of GABA, are also expressed in geniculate<br />

ganglia. Immunostaining shows that a majority of geniculate<br />

ganglion neurons express variable levels of GABAA-a1 in their<br />

somata. Fibers projecting peripherally from geniculate ganglia<br />

were strongly immunopositive <strong>for</strong> GABA-a1, a pattern very<br />

similar between GABAA-a1 and P2X2. In summary, GABA<br />

may function <strong>for</strong> both cell-to-cell communications within taste<br />

buds and in afferent taste neurotransmission. Functional imaging<br />

studies are in progress to test whether inhibitory responses to<br />

GABA are detected in all or a subset of geniculate ganglion<br />

neurons. Acknowledgements: NIH/NIDCD R01DC6308 and<br />

NIH/NIDCD R21DC12746<br />

#P152 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Expression and Signaling of IL-10 in Taste Cells<br />

Pu Feng, Jinghua Chai, Hong Wang<br />

Monell Chemical Senses Center Philadelphia, PA, USA<br />

Aged taste cells in taste buds are continuously replaced by<br />

young cells differentiated from basal cells. However, the related<br />

mechanisms that regulate survival and death of taste cells<br />

remain largely unknown. Here we continue to study the roles<br />

of cytokines tumor necrosis factor (TNF) and interleukin-10<br />

(IL-10), in taste cell turnover. TNF, a proinflammatory cytokine,<br />

often induces apoptosis by binding to its receptors and activating<br />

cell death pathways. In contrast, IL-10, an anti-inflammatory<br />

cytokine, suppresses TNF expression and promotes cell<br />

growth and survival. We previously found that T1R3 + type II<br />

taste cells are TNF-producing cells, and TNF receptors are<br />

globally expressed in taste buds. Here we report the expression<br />

and signaling of IL-10 in taste buds. We find that IL-10 and<br />

its receptor IL10R1 are preferentially expressed by different<br />

subsets of type II taste cells. Based on immuno-colocalization<br />

experiments using taste cell-type markers, the IL-10-producing<br />

cells are predominantly type II taste cells expressing the<br />

G-protein a-gustducin, while IL10R1 is selectively present in<br />

taste cells that express T1R3 and TNF. The IL-10 production<br />

can be induced by microbial products lipopolysaccharide (LPS)<br />

and Staphylococcal Enterotoxin A (SEA). The LPS-induced<br />

IL-10 expression in taste cells was profoundly diminished in<br />

TLR2-TLR4 double-gene-knockout mice, which indicates the<br />

dependence of IL-10 induction on TLR signaling pathways. The<br />

results strongly suggest that IL-10 produced by a-gustducin+<br />

type II cells could specifically down-regulate TNF production by<br />

acting on IL10R in T1R3+TNF+ type II cells. This interaction<br />

between the two subsets of taste cells may provide a mechanism<br />

<strong>for</strong> cellular crosstalk in taste buds under circumstances such as<br />

injury, infection and inflammation. Acknowledgements: The<br />

study was supported by NIH/NIDCD grants DC010012 and<br />

DC011735 and NSF grant DBJ-0216310<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

86


#P153 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Bitter Taste Stimuli trigger functionally distinct and<br />

interdependent Gustatory Signaling Pathways in<br />

immortalized Human Taste Cells<br />

Andreas Hochheimer, Michael Krohn, Holger Zinke<br />

B.R.A.I.N AG Zwingenberg, Germany<br />

Stably proliferating human taste cell lines are a powerful tool<br />

to gain new insights into human taste reception and signal<br />

transduction mechanisms. We previously established human<br />

taste cell lines from lingual epithelium, which maintain their<br />

endogenous programming and dedication to bitter taste<br />

reception. HTC-8 cells express 15 of 25 human TAS2R bitter<br />

taste receptor genes and respond to various bitter stimuli with<br />

endogenous gustatory signaling including calcium signaling and<br />

cell membrane depolarization. We used Fluo-4 calcium imaging<br />

assays and FLIPR fluorescent membrane potential assays to<br />

measure responses of human taste cells to various bitter taste<br />

stimuli and combinations of bitter taste stimuli. Our results<br />

revealed that stimulation with salicin elicits a PLC-dependent<br />

increase of intracellular calcium from internal calcium stores<br />

and does not lead to membrane depolarization. In contrast,<br />

addition of other bitter taste stimuli <strong>for</strong> instance PTC, saccharin<br />

and aristolochic acid led to cell membrane depolarization as<br />

well as to PLC-independent increase of intracellular calcium,<br />

which depends on extracellular calcium. These results suggest<br />

that gustatory responses to bitter stimuli are not uni<strong>for</strong>m in<br />

human taste cells and that bitter taste stimuli may trigger distinct<br />

signaling pathways. To test, whether these distinct signaling<br />

pathways interact we stimulated HTC-8 cells with salicin in<br />

combination with PTC, saccharin or aristolochic acid in the<br />

absence of extracellular calcium. Surprisingly, even though PTC,<br />

sacharin and aristolochic acid alone elicited no response, the<br />

PLC-dependent increase of intracellular calcium in response to<br />

salicin was strongly enhanced. These results suggest that crosstalk<br />

between bitter taste receptors and/or signaling pathways<br />

may occur in human taste cells. Acknowledgements: The<br />

research was funded by BRAIN AG corporate funds.<br />

#P154 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Sodium-sensing cells in mouse vallate taste buds<br />

Anthony Huang 1,3 , Stephen D Roper 1,2<br />

1<br />

Department of Physiology & Biophysics, University of Miami<br />

Miller School of Medicine Miami, FL, USA, 2 Program in Neuroscience,<br />

University of Miami Miller School of Medicine Miami, FL, USA,<br />

3<br />

Department of Anatomy, Southern Illinois University Carbondale,<br />

IL, USA<br />

Taste buds contain two types of cells that directly participate<br />

in gustatory transduction-- Receptor (Type II) and Presynaptic<br />

(Type III) cells. Receptor cells respond to sweet, bitter and<br />

umami taste stimulation. Presynaptic cells have been identified<br />

as sour (acid) taste-sensing cells. Using confocal Ca 2+ imaging<br />

in a lingual slice preparation, Tomchik et al. (2007) showed that<br />

some vallate taste cells responded to salt (Na + ) taste. Recently,<br />

ion channels (ENaCs) believed to underlie salt taste transduction<br />

were reported to be in taste cells that were neither Receptor<br />

(Type II) nor Presynaptic (Type III) cells. Yet, mechanisms<br />

of Na + taste transduction and the identity of the cells that<br />

respond to Na + stimulation remain controversial. Using Ca 2+<br />

imaging on isolated mouse vallate taste cells loaded with Fura<br />

2, we demonstrate here that a subset of cells, neither Receptor<br />

nor Presynaptic cells, show Ca 2+ mobilization in response to<br />

Na + stimulation. Removing extracellular Ca 2+ had no effect<br />

on responses evoked by Na + stimulation. In marked contrast,<br />

applying 1 μM thapsigargin, a SER Ca-ATPase inhibitor, or<br />

10 μM U73122, a PLC blocker, eliminated Na + responses,<br />

suggesting that Na + stimulation triggers PLC/IP3-mediated<br />

release of Ca 2+ from intracellular stores. Next, we used CHO<br />

cells expressing P2X receptors as biosensors to monitor ATP<br />

release from taste buds and isolated taste cells. Na + stimulation<br />

elicited robust biosensor responses that were blocked by suramin,<br />

confirming that Na + stimulation elicits ATP secretion from taste<br />

buds/cells. Carbenoxolone (5 μM) or probenecid (250 μM),<br />

blockers of pannexin1 hemichannels, reversibly blocked Na + -<br />

evoked ATP secretion. Our data indicate that salt taste stimulus<br />

triggers a dedicated population of taste cells to secrete ATP via<br />

pannexin1hemmichannels. Acknowledgements: Supported by<br />

NIH/NIDCD 5R01DC007630 (SDR).<br />

#P155 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

IL-1b Enhances Sodium Transport in Taste Buds<br />

D Kumarhia<br />

Institute of Molecular Medicine and Genetics Augusta, GA, USA<br />

Sodium ions pass through apical epithelial sodium channels<br />

(ENaC) in taste cells, depolarizing them and transmitting<br />

in<strong>for</strong>mation about sodium taste to the brain. Salt taste<br />

transduction is among the least understood of the taste<br />

transduction pathways. Moreover, few ENaC modulators have<br />

been identified in taste receptor cells compared to epithelial cells<br />

from kidney, lung and colon. Interleukin (IL)-1b is a classical<br />

proinflammatory cytokine produced by activated leukocytes.<br />

IL-1b and its receptor are also strongly expressed in taste cells.<br />

This cytokine promotes the maintenance of normal sodium taste<br />

function after contralateral chorda tympani injury, but the direct<br />

effects of IL-1b on sodium transport in taste buds are unknown.<br />

We loaded rat lingual epithelia containing fungi<strong>for</strong>m taste buds<br />

with the sodium indicator dye CoroNa Green, and measured<br />

changes in fluorescence (F 490<br />

) in response to basolateral IL-1b.<br />

Apical administration of the ENaC blocker, amiloride (50 µM in<br />

control Ringer’s solution), decreased F 490<br />

by 10-15%. Basolateral<br />

IL-1b (0.05 ng/ml – 5 ng/ml in control Ringer’s solution) caused<br />

upsurges in F 490<br />

resulting in an overall 2-10% increase in sodium<br />

transport above baseline. The effects of IL-1b were at least<br />

partially amiloride-sensitive, and occurred within seconds.<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

87


These results indicate that IL-1b rapidly augments ENaC<br />

function, in contrast to the cytokine’s delayed inhibition<br />

of sodium transport in other epithelial cells. Together, this<br />

indicates a novel, direct influence of IL-1b on ENaC in taste<br />

buds, mediated by mechanisms which diverge from those acting<br />

in other sodium-transporting epithelia. Acknowledgements:<br />

NIDCD 005811-10<br />

#P157 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Ryanodine receptors selectively interact with L type calcium<br />

channels in mouse taste cells<br />

Amanda B Maliphol 1 , Michelle R Rebello 1,2 , Kathryn F Medler 1<br />

1<br />

University at Buffalo Buffalo, NY, USA, 2 John B Pierce Lab New<br />

Haven, CT, USA<br />

#P156 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Ligand Specificity of Orphan G Protein-coupled<br />

Receptor GPR84<br />

Yan Liu, Timothy A. Gilbertson<br />

Department of Biology, Utah State University Logan, UT, USA<br />

Previous studies concluded that medium chain fatty acids<br />

with carbon chain lengths of 9-14 were ligands <strong>for</strong> GPR84<br />

(Venkataraman and Kuo Immunol Lett 101:144, 2005; Wang et<br />

al. J Biol Chem 281:34457, 2006), however, there has never been<br />

a careful and systematic analysis of the ligand specificity of this<br />

receptor which we have shown previously to be expressed in<br />

mammalian taste cells. As a means to compare the specificity<br />

and concentration-response functions <strong>for</strong> fatty acids in the taste<br />

system with GPR84, fura-2 based ratiometric calcium imaging<br />

was used to characterize GPR84 in a cell line that has been<br />

designed to express this receptor in an inducible fashion under<br />

control by the tetracycline (TET) promoter. The specific cell line<br />

has an inducible GPR84 + Gqi9 (a chimeric G protein; (Wang<br />

et al., 2006)), which has been cloned, validated by PCR/qPCR<br />

and verified <strong>for</strong> function in FLIPR-based calcium assays. Noninduced<br />

cells were used as controls. Using fura-2 based calcium<br />

imaging, we found that caproic (C 6:0<br />

), caprylic acid (C 8:0<br />

), capric<br />

acid (C 10:0<br />

), undecanoic acid (C 11:0<br />

), lauric acid (C 12:0<br />

), oleic acid<br />

(C 18:1<br />

), and arachidic acid (C 20:0<br />

) can all elicited a robust and<br />

reversible increase in intracellular calcium in the cells induced<br />

to express GPR84, while they cannot induce any calcium signal<br />

in the non-induced cells. Our results suggest GPR84 functions<br />

as a receptor <strong>for</strong> unsaturated fatty acids from C 6:0<br />

–C 12:0<br />

and<br />

other fatty acids (oleic acid, arachidic acid) previously not<br />

thought to activate this receptor. Currently, we are investigating<br />

the concentration-response functions <strong>for</strong> ligands of GPR84 in<br />

the inducible cell line. The specific signaling pathway <strong>for</strong> fatty<br />

acid transduction through GPR84 receptors and its role in the<br />

peripheral gustatory system remain to be elucidated.<br />

We reported that ryanodine receptors, specifically ryanodine<br />

receptor type 1, are expressed in two different types of<br />

mammalian peripheral taste receptor cells: Type II and Type<br />

III cells. In Type II cells that lack voltage-gated calcium<br />

channels (VGCCs) and chemical synapses, the ryanodine<br />

receptors contributed to the taste-evoked calcium signals that are<br />

initiated by opening inositol trisphosphate receptors located on<br />

internal calcium stores. In Type III cells that do have VGCCs<br />

and chemical synapses, ryanodine rreceptors were no longer<br />

able to contribute to taste-evoked calcium release signals but<br />

contributed to the depolarization-dependent calcium influx.<br />

The goal of this study was to better understand the role of the<br />

ryanodine rreceptors in Type III cells. Specifically, we wished<br />

to establish if there was selectivity in the type of VGCC that<br />

was associated with the ryanodine receptor or if the ryanodine<br />

receptor opened irrespective of the calcium channels involved.<br />

We also wished to determine if the ryanodine receptors and<br />

VGCCs required a physical linkage to interact or were simply<br />

functionally associated with each other. Using calcium imaging<br />

and pharmacological inhibitors on a transgenic mouse line<br />

that expresses green fluorescent protein (GFP) in GAD67<br />

expressing Type III taste cells, we found that ryanodine receptors<br />

are selectively associated with L type VGCCs but not through<br />

a physical linkage. Taste cells are able to undergo calcium<br />

induced calcium release through ryanodine receptors to increase<br />

the initial calcium influx signal and provide a larger calcium<br />

response than would otherwise occur when L type channels are<br />

activated. Acknowledgements: This work was supported by NSF<br />

Grant 0917893 to KFM.<br />

#P158 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Functional Profile of the Adult Glossopharyngeal Nerve<br />

Following Neonatal Chorda Tympani Transection in Rats<br />

Louis J. Martin, Suzanne I. Sollars<br />

University of Nebraska at Omaha/Psychology Department Omaha,<br />

NE, USA<br />

Rats receiving bilateral neonatal chorda tympani transection<br />

(neoCTX) show an increased preference <strong>for</strong> ammonium<br />

chloride (NH 4<br />

Cl) as adults – a substance which normal adult<br />

rats never prefer. It is currently unclear what changes to the<br />

taste system underlie this altered preference. To determine if<br />

injury-induced differences in the response properties of the<br />

remaining taste nerves can account <strong>for</strong> this behavior, whole-nerve<br />

electrophysiology was per<strong>for</strong>med on the glossopharyngeal nerve<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

88


(GL) of adult rats that underwent either unilateral neoCTX or<br />

a control surgery at five days of age. Nerve activity following<br />

application of various concentrations of NH 4<br />

Cl, NaCl, and<br />

sucrose was recorded using 0.5M NH 4<br />

Cl and 0.5M KCl as<br />

standards. There were no differences in nerve response to NH 4<br />

Cl<br />

or NaCl, but there was a significant difference <strong>for</strong> sucrose, with<br />

neoCTX rats having higher GL responses to the tastant. Since<br />

NH 4<br />

Cl responses did not differ between surgical groups, there<br />

may be differences following neoCTX in greater superficial<br />

petrosal or superior laryngeal nerve activity, or alterations in<br />

central processing which can account <strong>for</strong> the increase in NH 4<br />

Cl<br />

preference. Alternatively, the functioning of the GL may be more<br />

affected following bilateral compared to unilateral neoCTX. The<br />

mechanisms which lead to the observed injury-induced alteration<br />

in sucrose responding are unknown. However, this increase in<br />

responding following neoCTX could help explain observations<br />

from the human literature in which early chorda tympani<br />

damage is correlated with an increased preference <strong>for</strong> sugary<br />

foods. Work is currently underway to record GL responses from<br />

adult rats receiving bilateral neoCTX at 10 days of age.<br />

#P159 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Mouse bitter taste<br />

Wolfgang Meyerhof 1 , Kristina Lossow 1 , Hübner Sandra 1 ,<br />

Frenzel Sabine 1 , Masataka Narukawa 1 , Anja Voigt 1 , Ulrich Boehm 2 ,<br />

Jonas Töle 1 , Maik Behrens 1<br />

1<br />

German Institute of Human Nutrition Potsdam-Rehbruecke,<br />

Department of Molecular Genetics Nuthetal, Germany, 2 University<br />

of Saarland, School of Medicine, Department of Pharmacology and<br />

Toxicology Homburg, Germany<br />

extents. Behavioral experiments showed that these mice exhibit<br />

diminished avoidance of several bitter compounds, whereas they<br />

are indistinguishable from control mice in their avoidance of<br />

denatonium benzoate. Together the data demonstrate that bitter<br />

sensing cells are functionally polymorphic <strong>for</strong>ming the basis <strong>for</strong><br />

variable behavioral responses to different bitter chemicals.<br />

#P160 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Behavioral responses to trimethylamine -N -oxide using<br />

the CTA paradigm in C57BL/6 mice<br />

Yuko Murata, Meiko Kimura<br />

Fisheries Research Agency Yokohama, Japan<br />

Trimethylamine-N-oxide (TMAO) is a common and compatible<br />

osmolyte in tissues of marine organisms, and counteracts<br />

the effects of protein destabilizing agents such as urea in<br />

elasmobranches. Gadoid fish, elasmobranches and scallop have<br />

high levels of TMAO in their muscles. TMAO is thought to<br />

contribute to the taste of these fishes but the taste of TMAO is<br />

unclear. In this study, we investigated taste quality perception<br />

of TMAO in C57BL/6 mice using conditioned taste aversion<br />

(CTA) experiments.We developed LiCl-induced CTA to 1.0%<br />

TMAO and examined its generalization to 12 taste stimuli.<br />

CTA to TMAO significantly generalized to D-phenylalanine,<br />

saccharine and quinine. These results suggest that mice avoid<br />

the taste of TMAO and its taste perception in mice is similar<br />

to D-phenylalanine, saccharin and quinine. We will also<br />

present behavioral thresholds and behavioral response to low<br />

concentration (below 0.1%) of TMAO.<br />

Depending on dose, bitter chemicals can be toxic or healthy.<br />

Accordingly, consumers like some bitter tasting foods while<br />

they avoid others. For a detailed understanding of bitter taste<br />

physiology we examined the receptors <strong>for</strong> bitter substances and<br />

their cells in mice. Functional expression analysis showed that<br />

mice and humans detect a similar range of bitter compounds<br />

even though mice possess 30% more Tas2r bitter taste receptor<br />

genes than humans. Intriguingly, recognition of bitter chemicals<br />

in the two species is mostly evoked by non-orthologous Tas2rs.<br />

Based on the number of cognate compounds, mouse Tas2rs,<br />

like their human counterparts, can be classified in generalists,<br />

moderately tuned receptors and specialists. However, compared<br />

with humans, mice seem to possess a larger fraction of specialists<br />

suggesting that a greater number of Tas2r genes offers the luxury<br />

of a set of specific receptors <strong>for</strong> selected bitter compounds.<br />

Genetic labeling and in situ hybridization experiments revealed<br />

that mice possess 2200 to 3300 bitter-sensing cells. They express<br />

the entire repertoire of Tas2r genes at individual levels and<br />

in limited subsets. Accordingly, genetic ablation of the cells<br />

expressing one Tas2r, i.e., Tas2r131, did not extinguish the entire<br />

population of bitter sensing cells but only ~50%. The remaining<br />

bitter sensing cells displayed complete absence of some<br />

Tas2rs, whereas expression of others was reduced to different<br />

#P161 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Purification, biophysical characterization and first<br />

crystallization trials of the ligand-binding domain<br />

of the human T1R3 sweet taste receptor.<br />

Fabrice Neiers, Maud Sigoillot, Elodie Maîtrepierre, Christine Belloir,<br />

Nicolas Poirier, Loïc Briand<br />

Centre des <strong>Sciences</strong> du Goût et de l’Alimentation, INRA CNRS<br />

Université de Bourgogne Dijon, France<br />

The heterodimeric T1R2/T1R3 sweet taste receptor is<br />

composed of two class C G-protein coupled receptors<br />

(GPCRs), while T1R1/T1R3 heterodimer is involved in umami<br />

taste perception. Class C GPCRs share common structural<br />

homologies including a large N-terminal domain (NTD) linked<br />

to the seven transmembrane domain by a cysteine rich region.<br />

T1R2- and T1R1-NTDs have been shown to contain the primary<br />

binding site <strong>for</strong> most of the sweet ligands and umami tastants,<br />

respectively. In contrast, the contribution of T1R3-NTD to sweet<br />

and umami compound detection is less documented. The human<br />

T1R3-NTD was produced in Escherichia coli using a strategy<br />

recently described (Maîtrepierre et al., Protein Expr. Purif., 2011).<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

89


The purified protein was characterized by SDS–PAGE, circular<br />

dichroism and a fluorescent ligand-binding assay. Using size<br />

exclusion chromatography coupled to static light scattering,<br />

we showed that hT1R3-NTD is monodisperse and <strong>for</strong>ms<br />

homodimers. These data demonstrate that the purified protein is<br />

not only suitable <strong>for</strong> functional analyses, but also <strong>for</strong> subsequent<br />

crystallization trials. Acknowledgements: This work was<br />

supported by fundings from INRA, Burgundy council (Région<br />

Bourgogne) and Agence Nationale de la Recherche (ANR-09-<br />

ALIA-010).<br />

#P162 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Regulation of basal sweet sensitivity of mice by leptin.<br />

Mayu Niki, Masafumi Jyotaki, Tadahiro Ohkuri, Ryusuke Yoshida,<br />

Yuzo Ninomiya<br />

Section of Oral Neuroscience, Graduate School of Dental <strong>Sciences</strong>,<br />

Kyushu University Fukuoka, Japan<br />

Leptin (Lep) is shown to selectively suppress sweet taste<br />

responses in lean mice but not in Lep receptor-deficient db/<br />

db mice. In contrast, endocannabinoids (EDs) enhance sweet<br />

taste sensitivities in lean mice but not in mice genetically lacking<br />

CB 1<br />

receptors. However the action of endogenous Lep and<br />

EDs on taste responses is not fully understood. In this study, we<br />

examined expression of related molecules, the effect of leptin<br />

on taste cell responses and the effect of antagonists <strong>for</strong> Ob-Rb<br />

(leptin L39A/D40A/F41A : LA) and CB 1<br />

(AM251) on the<br />

chorda tympani (CT) responses in mice with different serum Lep<br />

levels. About 40 % of taste cells expressing T1r3 coexpressed<br />

Ob-Rb and a subset of taste cells expressed biosynthesizing<br />

enzyme (DAGL) and degrading enzyme (MAGL) of ED (2-<br />

AG). In about half of sweet sensitive taste cells, bath application<br />

of 20 ng/ml leptin suppressed responses to sweeteners (


the hyperpolarizing conductance and sustain the smooth fast<br />

swimming. We have used RNA interference (RNAi) to test<br />

several candidate genes <strong>for</strong> possible function as an L-glutamate<br />

receptor and found a gene with sequence homology to the<br />

glutamate binding subunit of an NMDA-like receptor. When<br />

the mRNA <strong>for</strong> this gene, pGluR1, is depleted using RNAi,<br />

the cells are not attracted to L-glutamate, but continue to be<br />

attracted to D-glutamate. To better understand the signaling<br />

pathway <strong>for</strong> L-glutamate, we have epitope tagged the pGluR1<br />

protein and found by Western blotting and mass spectrometry<br />

that it is in the membrane of the cell body and cilia. A catalytic<br />

subunit of the ciliary adenylyl-cyclase labeled with GFP does<br />

not co-immuniprecipitate with FLAG-pGluR1. Mammalian<br />

NMDA-like receptors are dependent upon glycine and D-serine,<br />

but the P. tetraurelia chemoresponse to L-glutamate is unaffected<br />

by combinations of glycine and D-serine and L-glutamate.<br />

Acknowledgements: NIH Grant Number 2 P20 RR016435-06 to<br />

the Neuroscience Center of Research Excellence (COBRE to R.<br />

Parsons, PI) <strong>for</strong> imaging support and NIH grant RO1 GM59988<br />

taste stimuli, including acids and salts, were abolished. Taste<br />

responses also are blocked in a dose-dependent fashion with<br />

application of higher concentrations of AF-353 directly to the<br />

tongue and can be partially recovered upon washout of the drug.<br />

These data clearly indicate that activation of P2X receptors and<br />

there<strong>for</strong>e ATP release is required <strong>for</strong> all taste modalities in mice.<br />

Acknowledgements: Funded by NIH grants R01 DC012555,<br />

R01 DC007495, and P30 DC04657 and a gift from Afferent<br />

Pharmaceuticals.<br />

#P166 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

Nonsynaptic Contacts in Rat Circumvallate Taste Buds:<br />

Subsurface Cisternae and Atypical Mitochondria<br />

Ruibiao Yang 1,2 , Amanda E. Bond 1,2 , John C. Kinnamon 1,2<br />

1<br />

Department of Biological <strong>Sciences</strong>, University of Denver Denver, CO,<br />

USA, 2 Rocky Mountain Taste and Smell Center Aurora, CO, USA<br />

#P165 POSTER SESSION III:<br />

TRIGEMINAL; HUMAN OLFACTORY<br />

PSYCHOPHYSICS; TASTE PERIPHERY<br />

A selective P2X3, P2X2/3 receptor antagonist abolishes<br />

responses to all taste stimuli in mice<br />

Aurelie Vandenbeuch 1,3 , Catherine B. Anderson 1,3 , Anthony P. Ford 4 ,<br />

Steve Smith 4 , Thomas E. Finger 2,3 , Sue C. Kinnamon 1,3<br />

1<br />

University of Colorado School of Medecine, Department of<br />

Otolaryngology Aurora, CO, USA, 2 University of Colorado School of<br />

medecine, Department of Cell and Development Biology Aurora, CO,<br />

USA, 3 Rocky Mountain Taste and Smell Center Aurora, CO, USA,<br />

4<br />

Afferent Pharmaceuticals San Mateo, CA, USA<br />

ATP is believed to be a crucial neurotransmitter <strong>for</strong><br />

communicating gustatory in<strong>for</strong>mation to nerve fibers. Evidence<br />

is based largely on recordings from mice lacking both P2X2<br />

and P2X3 purinergic receptor subunits (P2X2-P2X3 DKO<br />

mice), which lack responses to all taste stimuli (Finger et al.,<br />

2005). These data suggest that all taste qualities require ATP to<br />

communicate with nerve fibers. However, subsequent studies<br />

have detected ATP release only from Type II taste cells, those<br />

that respond to bitter, sweet, and umami stimuli (Huang et<br />

al., 2007; Romanov et al., 2007). Recent experiments on the<br />

P2X2-P2X3 DKO mice have shown that in addition to the lack<br />

of postsynaptic receptors, these mice fail to release ATP to taste<br />

stimuli (Huang et al., 2011). Thus, the lack of taste responses<br />

may be due to the lack of ATP release rather than the lack of<br />

postsynaptic receptors. To resolve whether postsynaptic P2X<br />

receptors are required <strong>for</strong> transmission of all tastes to the nervous<br />

system, we have used a pharmacological approach to chemically<br />

block the purinergic receptors while recording from the chorda<br />

tympani nerve in response to a battery of taste stimuli. C57Bl6<br />

mice were injected ip with 6 mg/kg AF-353, a membrane<br />

permeant compound that blocks all P2X3-containing receptors<br />

(P2X3 homotrimers and P2X2/3 heterotrimers; Gever et al.,<br />

2010). Within 15 min of injection, integrated responses to all<br />

It is generally accepted that Type II cells transduce bitter,<br />

sweet, and umami stimuli in rodent taste buds. Recent studies<br />

also demonstrate that Type II taste cells release ATP, a<br />

neurotransmitter believed to play an important role in taste<br />

transduction. However, no classical synapses have been found<br />

to be associated with Type II cells. Are there other contacts<br />

between Type II cells and Type III cells or Type II cells and<br />

nerve processes? In the present study, we utilized conventional<br />

transmission electron microscopy to examine the ultastructural<br />

features of nonsynaptic contacts in rat circumvallate taste buds.<br />

Our results indicate that Type II cells are in intimate contact with<br />

Type III cells in taste buds. Two types of nonsynaptic contacts,<br />

subsurface cisternae of endoplasmic reticulum and/or atypical<br />

mitochondria, have been found to be present adjacent to the<br />

cytoplasmic leaflet of Type II cells at close appositions with<br />

nerve processes. Frequently we observed subsurface cisternae of<br />

smooth or rough endoplasmic reticulum in Type II cells at sites<br />

of apposition with nerve processes. Occasionally we observed<br />

atypical mitochondria and subsurface cisternae adjacent to<br />

each other in Type II cells. We also observed electron-dense<br />

periodic pillars connecting the subsurface cisternae or the<br />

atypical mitochondria to the membranes of Type II cells at<br />

close appositions with nerve processes. We speculate that these<br />

structures are involved in communication between Type II cells<br />

and nerve processes. Acknowledgements: This work is supported<br />

by NIH grants DC00285 and P30 DC04657<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

91


#P167 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

#P168 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

A role <strong>for</strong> bitter taste receptors in thyroid toxicity and<br />

hormone production<br />

Adam A. Clark 1,2 , Cedrick D. Dotson 1 , Amanda E.T. Elson 1 ,<br />

Nanette I. Steinle 3 , Steven D. Munger 1,2,3<br />

1<br />

University of Maryland Baltimore Department of Anatomy and<br />

Neurobiology Baltimore, MD, USA, 2 University of Maryland Baltimore<br />

Program in Toxicology Baltimore, MD, USA, 3 University of Maryland<br />

Baltimore Department of Medicine Baltimore, MD, USA<br />

Bitterness is associated with toxicity. Bitter compounds activate<br />

a small family of G protein-coupled taste receptors (T2Rs)<br />

expressed in the taste buds. T2Rs are also found in non-gustatory<br />

tissues (e.g., the respiratory and gastrointestinal systems),<br />

suggesting that many ingested or inhaled toxins could have<br />

broad physiologic effects. We report that T2Rs are expressed<br />

in thyroid follicular cells (FCs), where they modulate iodide<br />

efflux. Immunohistochemical and PCR analyses show that<br />

numerous T2Rs as well as the T2R-associated G protein subunit<br />

a-gustducin are expressed in human and rodent FCs and in a<br />

human follicular cell line, Nthy-Ori 3-1. Thyroid stimulating<br />

hormone (TSH)-dependent Ca 2+ signals, an important regulator<br />

of iodide efflux from FCs, were significantly reduced in Nthy-<br />

Ori 3-1 cells by the T2R ligands denatonium benzoate (DB),<br />

chloramphenicol (Chlor) and cycloheximide (Cyx). By contrast,<br />

the T2R38 ligand 6-n-propylthiouracil (PROP) had no effect<br />

on TSH-dependent Ca 2+ signals, likely because this cell line<br />

expresses only the “non-taster” T2R38 variant. DB, Chlor and<br />

Cyx also significantly reduced TSH-dependent iodide efflux from<br />

Nthy-Ori 3-1 cells. Decreased iodide efflux in vivo should result in<br />

decreased production of the thyroid hormones triiodothyronine<br />

(T3) and thyroxine (T4). Consistent with this, we found that<br />

a nonsynonymous polymorphism in T2R42 is associated<br />

with lower free T3 and T4 levels in a human cohort. Thyroid<br />

hormones can affect energy expenditure, thermoregulation, body<br />

weight, and body composition. T2Rs may be a useful target <strong>for</strong><br />

pharmacologic regulation of these endocrine signals, perhaps<br />

leading to new interventions <strong>for</strong> chronic morbidities involving<br />

fatigue or obesity. Acknowledgements: Support: NIDCD (R01<br />

DC010110), NIDDK (P30 DK072488).<br />

Metabolic effects of long-term sweetener consumption<br />

Maartje CP Geraedts 1 , Tatsuyuki Takahashi 1 , Stephan Vigues 1 ,<br />

Cedric Uytingco 1 , Steven D Munger 1,2<br />

1<br />

University of Maryland School of Medicine, Department of Anatomy<br />

& Neurobiology Baltimore, MD, USA, 2 University of Maryland School<br />

of Medicine, Department of Medicine, Division of Endocrinology,<br />

Diabetes and Nutrition Baltimore, MD, USA<br />

The sweet taste receptor T1R2+T1R3 responds to diverse<br />

gustatory stimuli including sugars and low calorie sweeteners<br />

(LCS). In intestinal enteroendocrine L cells, T1R2+T1R3<br />

couples glucose stimulation to the secretion of insulinotropic<br />

hormone glucagon-like peptide-1 (GLP-1). While the<br />

preponderance of in vivo evidence indicates that LCS do not<br />

impact glucose homeostasis, the responsiveness of T1R2+T1R3<br />

to natural and artificial LCS suggests that these compounds<br />

could exert extraoral effects. To address this issue, we have<br />

initiated long-term sweetener (300 mM sucrose, Su; 1 mM<br />

sucralose, Sa; 10 mM cyclamate, Cy; 3 mM acesulfame K; 5 mM<br />

rebaudioside A) consumption studies in mice with (T1R3 +/+ ) or<br />

without (T1R3 -/- ) a functional sweet taste receptor. Mice (4 w.o.)<br />

are maintained <strong>for</strong> 3, 6, 9 or 12 mo. on normal chow diets along<br />

with ad lib water containing one of the sweeteners. Cy serves as<br />

a negative control as it is not an agonist <strong>for</strong> mouse T1R2+T1R3.<br />

Mice are assessed <strong>for</strong> food and fluid intake, sweet taste<br />

preference, body weight, body fat, metrics of glucose and insulin<br />

homeostasis, and glucose-stimulated GLP-1 secretion from ileum<br />

and colon explants. Initial results indicate that both T1R3 +/+<br />

and -/- mice maintained on 300mM Su have significantly higher<br />

body fat content than either Sa- or Cy-treated mice. By contrast,<br />

Sa-treated mice show a strong potentiation of glucose-stimulated<br />

GLP-1 secretion (as compared to Su- or Cy-treated mice) at 3<br />

mo, with reduced potentiation at later time points. Interestingly,<br />

glucose tolerance tests revealed no significant differences<br />

in glucose or insulin levels across these sweetener groups.<br />

Ongoing assessments of all five sweeteners will be presented.<br />

Acknowledgements: Tate&Lyle and NIDCD (DC010110)<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

92


#P169 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

#P170 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

Oral stimulation with sugar elicits cephalic phase insulin<br />

release and improves glucose tolerance in C57BL/6 and<br />

Tas1r3 knock-out mice<br />

John I. Glendinning 1 , Sarah Stano 1 , Olivia Goldman 1 , Danica Yang 1 ,<br />

Robert F. Margolskee 2 , Anthony Sclafani 3 , Joseph R. Vasselli 4<br />

1<br />

Barnard College, Columbia University/Biology Department New<br />

York, NY, USA, 2 Monell Chemical Senses Center Philadelphia, PA,<br />

USA, 3 Brooklyn College of CUNY/Psychology Department Brooklyn,<br />

NY, USA, 4 New York Obesity Nutrition Research Center, St. Luke’s-<br />

Roosevelt Hospital New York, NY, USA<br />

In rats, oral stimulation by sugars elicits cephalic phase insulin<br />

release (CPIR). This CPIR has been found to improve tolerance<br />

to glucose challenges. Here, we asked whether oral stimulation<br />

by glucose has similar effects in C57BL/6 wild-type (WT) mice.<br />

To explore the role of the T1r3 subunit of the heterodimeric<br />

sweet taste receptor (T1r2+T1r3) in this process, we also<br />

examined Tas1r3 knock-out (KO) mice. We subjected all mice<br />

to a 2 g/kg glucose challenge. To this end, we administered a<br />

2.8 M glucose solution either orally (by allowing mice to take<br />

a predetermined number of licks) or post-orally (by injecting<br />

a specific volume into the esophagus by feeding tube). We<br />

collected tail blood immediately prior to glucose administration<br />

(at 0 min), and at subsequent time intervals. In Experiment<br />

1, we measured plasma insulin levels with an ELISA test. We<br />

inferred a CPIR if plasma insulin increased significantly between<br />

the 0 and 5 min measurements. Both strains of mice exhibited<br />

a significant CPIR after oral stimulation, but not after post-oral<br />

stimulation with glucose. The fact that both strains exhibited a<br />

CPIR of similar magnitude indicates that T1r3 is not necessary<br />

<strong>for</strong> eliciting the CPIR. In Experiment 2, we measured blood<br />

glucose levels at 0, 15, 30, 60 and 120 min. We found that both<br />

strains exhibited significantly better glucose tolerance when the<br />

glucose was administered orally than post-orally; the benefit<br />

of oral administration was especially pronounced in the Tas1r3<br />

KO mice. These results indicate that oral stimulation by sugars<br />

elicits a CPIR via a T1r3-independent taste mechanism, and that<br />

this CPIR substantially improves the ability of mice to tolerate<br />

glucose challenges. Acknowledgements: Summer Undergraduate<br />

Research Fellowship Program, Columbia University<br />

As American as Apple Pie Gum: A Study of Satiety<br />

Jack Hirsch 1 , Michele Soto 2 , Alan Hirsch 2<br />

1<br />

Adlai E. Stevenson High School Lincolnshire, IL, USA,<br />

2<br />

Smell & Taste Treatment and Research Foundation Chicago, IL, USA<br />

Purpose: The aim of this study is to determine the satiety<br />

value of Wrigley’s Extra Dessert Delight Apple Pie Sugar-Free<br />

Gum as compared to a 100 calorie slice of apple pie. Procedure:<br />

Thirteen girls and 11 boys self-assessed their degree of satiety<br />

on a visual analog scale be<strong>for</strong>e and after chewing Wrigley’s<br />

Extra Dessert Delight Apple Pie Sugar-Free Gum <strong>for</strong> one, 15,<br />

and 30 minutes. This was repeated on a separate day with a 100<br />

calorie slice of Market Pantry Apple Pie (Target) or vice-versa<br />

(the order of presentation being counter balanced). A satiety<br />

value was computed <strong>for</strong> each and statistical significance was<br />

determined <strong>for</strong> difference in satiety index between the two and<br />

effect of order of presentation (pie versus gum first). Conclusion:<br />

No statistically significant difference was seen in satiety value<br />

in response to eating a slice of apple pie or chewing apple<br />

pie-flavored gum (p=0.096). No effect of order presentation was<br />

seen. The gum imbued the same satiety value as apple pie, with<br />

less than 1/20 of the calories. The results suggest that chewing<br />

Wrigley’s Extra Dessert Delight Apple Pie Sugar-Free Gum may<br />

have a role in the promotion of satiety in children as part of a<br />

weight loss program.<br />

#P171 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

Pre-absorptive insulin release to glutamate taste<br />

Matthew C Kochem 1 , Suzanne M Alarcon 1 , Paul AS Breslin 1,2<br />

1<br />

Rutgers University New Brunswick, NJ, USA, 2 Monell Chemical<br />

Senses Center Philadelphia, PA, USA<br />

Glutamate, like glucose, requires insulin to be transported out<br />

of blood into tissue. Glutamate is detected in the mouth by<br />

taste cells, in the intestine by L-cells and in the pancreas by Beta<br />

cells. In addition, glutamate stimulates glucagon release from<br />

pancreatic alpha cells. It is unknown, however, whether preabsorptive<br />

insulin release (PIR) occurs in response to the taste of<br />

glutamate. The purpose of this study was to determine whether<br />

glutamate causes PIR and whether the magnitude of PIR was<br />

related to glutamate taste sensitivity, which is variable among<br />

individuals. To assess PIR, human subjects tasted and consumed<br />

a glutamate test meal at a dose of 100mg/kg body weight. The<br />

glutamate dose was comprised of a mixture of monosodium<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

93


glutamate and monopotassium glutamate. Subjects tasted the<br />

mixture by repeated swishing and spitting <strong>for</strong> 5 minutes. Subjects<br />

then ingested another dose of glutamate. Baseline blood samples<br />

were collected every 5 minutes be<strong>for</strong>e ingestion of glutamate.<br />

After ingestion, samples were collected at 3 minute intervals<br />

<strong>for</strong> a total of 15 minutes and analyzed <strong>for</strong> glucose, insulin and<br />

c-peptide. To assess glutamate sensitivity, subjects tasted and<br />

rated intensity matched NaCl, sucrose, and glutamate solutions.<br />

Relative glutamate sensitivity was defined as the ratio of<br />

perceived glutamate intensity to that of NaCl and sucrose. All<br />

seven subjects showed pre-absorptive blood glucose changes in<br />

response to the taste of glutamate. Six subjects showed increases<br />

in blood glucose (presumably due to glucagon release) and one<br />

subject showed a decrease. Half of the subjects showed preabsorptive<br />

changes in insulin. The highest PIRs occurred in the<br />

two subjects who showed the highest sensitivity to glutamate.<br />

The lowest PIR occurred in the subject with the lowest sensitivity<br />

to glutamate. Acknowledgements: NIH DC02995<br />

the control of dietary fat intake and that these effects appear<br />

to be significantly influenced by gender. Acknowledgements:<br />

Supported by NIH grant R01DK059611 (TAG)<br />

#P173 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

Human Anticipatory Blood Pressure Responses to Oral<br />

NaCl and KCl are Different<br />

Melissa A Murphy 1 , Paul A.S. Breslin 1,2<br />

1<br />

Department of Nutritional <strong>Sciences</strong>, Rutgers University New<br />

Brunswick, NJ, USA, 2 Monell Chemical Senses Center Philadelphia,<br />

PA, USA<br />

#P172 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

Mice lacking Trpm5 show reduced dietary fat intake<br />

Dulce M. Minaya, Jacqueline Lee, Dane R. Hansen,<br />

Timothy A. Gilbertson<br />

Utah State University/Biology Department Logan, UT, USA<br />

We recently showed a critical role of Trpm5 in the transduction<br />

pathway <strong>for</strong> long chain polyunsaturated fatty acids (Liu et al., J<br />

Neurosci 31: 8634, 2011). In the present study, we have begun to<br />

investigate dietary fat preference and the propensity to develop<br />

dietary-induced obesity in mice lacking Trpm5. In male mice<br />

placed on a high fat diet, those mice lacking Trpm5 (Trpm5 -/- )<br />

ate significantly less and accordingly weighed less and had less<br />

body fat than wild type (Trpm5 +/+ ) mice; no differences between<br />

Trpm5 -/- and Trpm5 +/+ mice were seen on a control (low fat) diet.<br />

Similar differences were recorded in control male mice and those<br />

lacking IP 3<br />

R 3<br />

, receptors that are upstream of Trpm5 activation in<br />

the fatty acid transduction pathway. Most surprisingly, however,<br />

was the fact that female mice with or without IP 3<br />

R 3<br />

did not<br />

show the same differences indicating a potential gender effect<br />

of this pathway on dietary fat intake. Given these data, we have<br />

per<strong>for</strong>med feeding studies on female Trpm5 -/- and Trpm5 +/+<br />

mice to look <strong>for</strong> similar gender-specific effects on fat intake.<br />

Like male mice, female mice lacking Trpm5 show a decrease in<br />

dietary fat intake and gain less weight than wild types, though<br />

these effects are much less robust and have a much slower onset<br />

than <strong>for</strong> the male mice. However, unlike the males, there is no<br />

significant difference in body fat between female Trpm5 -/- and<br />

Trpm5 +/+ after being on the high fat diet <strong>for</strong> 56 days. We are<br />

currently exploring the specificity of these effects <strong>for</strong> the different<br />

classes of fat (saturated versus unsaturated). Together, our data<br />

indicate that Trpm5 may play a role, directly or indirectly, in<br />

The mechanism of association between dietary sodium<br />

and blood pressure is not clearly defined, but high dietary<br />

salt intake is considered a risk <strong>for</strong> hypertension (HTN) and<br />

cardiovascular disease (CVD). Many studies have examined<br />

the impact of different dietary cations on these risk factors<br />

and implicate sodium (or Na/K ratio) as a key determinant.<br />

We previously presented data showing an anticipatory blood<br />

pressure (BP) response to the oral presentation of 1.0 Molar<br />

NaCl and suggested this may be linked to blood volume preabsorptive<br />

reflexes. Based on this premise, we hypothesized<br />

that this anticipatory reflex would be more pronounced <strong>for</strong><br />

sodium ions. To determine the specificity of the response to oral<br />

sodium, subjects rinsed orally with either 0.5 M KCl (matched<br />

<strong>for</strong> taste intensity to NaCl) to test cation composition and 1.0<br />

M Na-Gluconate to test salt taste intensity. Subjects rested<br />

in a seated position <strong>for</strong> 2.5 hours while we recorded resting<br />

BP and additional readings following the rinse at 10 minute<br />

intervals with a manual sphygmomanometer, one trial per day,<br />

five trials <strong>for</strong> each solution tested. In subjects whose blood<br />

pressure dropped following a rinse with NaCl, a similar trend<br />

was observed after a rinse with Na-Gluconate. Yet BP remained<br />

level, without a decreasing trend, following a rinse with KCl.<br />

Interestingly, the BP response to Na-Gluconate was weaker than<br />

the NaCl response, suggesting an influence of taste perception<br />

on the anticipatory BP reflex. The difference between the BP<br />

response to NaCl and to KCl suggests that there may be oral<br />

chemosensory specificity to the reflex. These data support the<br />

idea that anticipatory BP reflexes are cation specific and may<br />

involve gustatory mechanisms. Acknowledgements: Funded in<br />

part by NIH DC 02995 to PASB.<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

94


#P174 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

#P175 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

Odor Identification Per<strong>for</strong>mance in Middle Aged Obese<br />

Individuals with High Blood Pressure<br />

Stephanie M. Oleson 1 , Erin Green 3 , Aaron Jacobson 3 , Claire Murphy 1,2,3<br />

1<br />

San Diego State University San Diego, CA, USA, 2 University of<br />

Cali<strong>for</strong>nia, San Diego San Diego, CA, USA, 3 SDSU-UCSD Joint<br />

Doctoral Program in Clinical Psychology San Diego, CA, USA<br />

Obesity is a major health epidemic that affects people globally<br />

and is associated with the development of serious comorbidities<br />

such as diabetes, hypertension, and heart disease. The presence<br />

of obesity has also been linked to the risk <strong>for</strong> later development<br />

of Alzheimer’s Disease (AD) or mild cognitive impairment<br />

(MCI) and has also been associated with poorer per<strong>for</strong>mance<br />

on cognitive measures of global, executive, and memory<br />

functioning. Furthermore, olfactory functioning is linked to<br />

energy balance and metabolism and has been shown to be<br />

altered in obese individuals, as well as those diagnosed with<br />

AD or MCI. It has been proposed that the impact of obesity on<br />

cognition is indirect and influenced by comorbidities such as<br />

hypertension and diabetes. There<strong>for</strong>e, the current study sought<br />

to determine if odor identification per<strong>for</strong>mance differed between<br />

obese individuals with and without high blood pressure. Thirtyone<br />

obese individuals (BMI > 30 kg/m 2 ) between the ages of<br />

46-54 years old were given the San Diego Odor Identification<br />

test. Obese individuals with high blood pressure per<strong>for</strong>med<br />

significantly worse on the odor identification task ( F= 8.384; p<br />

= .008), but neither group significantly differed in BMI or odor<br />

threshold. The results suggest that obese individuals with high<br />

blood pressure are more likely to show odor memory deficits<br />

and that these changes occur as early as during middle-age.<br />

Acknowledgements: Supported by NIH grant #AG004085-25<br />

to CM.<br />

Effect of Oral Sensations on the Relief of Thirst<br />

Catherine Peyrot des Gachons 1 , Julie Avrillier 2 , Laure Algarra 3 ,<br />

Emi Mura 4 , Paul A.S. Breslin 1,5<br />

1<br />

Monell Chemical Senses Center Philadelphia, PA, USA,<br />

2<br />

AgroSup Dijon Institut National Superieur Dijon, France,<br />

3<br />

AgroParisTech Paris, France, 4 Suntory Business Expert Ltd.<br />

Kawasaki, Japan, 5 Rutgers University Department of Nutritional<br />

<strong>Sciences</strong> New Brunswick, NJ, USA<br />

Thirst is the internal sensation of a need to drink, presumably<br />

to rehydrate or recover bodily fluid losses. But thirst is quenched<br />

long be<strong>for</strong>e all ingested fluids are absorbed. There<strong>for</strong>e, sensory<br />

feedback must play a role in thirst quenching. Different beverages<br />

seem to quench thirst with different efficiencies, but how their<br />

oral sensory characteristics determine the thirst quenching<br />

efficacy is poorly understood. The purpose of this study was to<br />

determine which oral sensation(s) commonly manipulated in<br />

beverages, such as temperature or carbonation, influence levels<br />

of thirst. To answer this question, subjects who were deprived<br />

of liquid overnight were first asked to drink a fixed volume of<br />

an experimental beverage presenting one or two specific traits.<br />

Then we objectively evaluated their residual thirst by measuring<br />

how much additional plain, uncarbonated, room temperature<br />

water they wanted to drink afterwards. The results show that<br />

the perception of coldness is an important parameter <strong>for</strong> thirst<br />

quenching. A beverage at low temperature (5°C) quenches thirst<br />

more than a beverage at room temperature (20°C). Moreover,<br />

a cold, carbonated beverage relieves thirst even more than does<br />

a cold uncarbonated beverage. These results support, in part,<br />

the observations of the sensory controls of thirst quenching in<br />

the animal literature. Acknowledgements: Suntory Business<br />

Expert Ltd.<br />

#P176 PPOSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

Intraduodenal infusions of sucrose influence conditioned and<br />

unconditioned affective taste-guided responses to oral sucrose<br />

Lindsey A. Schier, Alan C. Spector<br />

Department of Psychology and Program in Neuroscience,<br />

Florida State University Tallahassee, FL, USA<br />

POSTER PRESENTATIONS<br />

Sensory signals ascending from the oral cavity and viscera are<br />

integrated in the CNS to adjust meal size and taste preferences.<br />

Electrophysiological data suggest such integrations affect early<br />

processing in brainstem taste nuclei. Taste receptors (e.g. T1R)<br />

are also found in GI cells, but whether taste-like signals arising<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

95


from postoral sites are integrated with oral taste-guided behaviors<br />

is unknown. This study examined effects of intraduodenal (ID)<br />

infusions of a sweet ligand on conditioned and unconditioned<br />

affective responses to matching oral chemosensory stimuli.<br />

First, rats were given two separate 15 min sessions to ingest<br />

0.3M sucrose directly followed by either ip LiCl (3 mEq/<br />

kg) to condition a sucrose taste aversion (CTA, n=9) or saline<br />

(unconditioned, n=8). Then, licking responses to 5 sucrose<br />

concentrations (0.01, 0.03, 0.1, 0.3, 1M), 0.12M NaCl, and water<br />

(10s trials in randomized blocks) were examined in two 30 min<br />

brief-access tests. Four min be<strong>for</strong>e each test, rats were infused<br />

ID with 0.3M sucrose or 0.15M NaCl (3ml). For unconditioned<br />

rats, ID sucrose enhanced preferential licking to 0.03-1 M oral<br />

sucrose, with no effect on licking <strong>for</strong> NaCl. This preference<br />

shift emerged rapidly by trial block 2. CTA rats reduced licking<br />

to 0.03-1M sucrose, not NaCl, but because CTA rats initiated<br />

so few trials, and significantly fewer after ID sucrose versus ID<br />

NaCl, detection of emergent differences due to ID preload type<br />

was likely limited. Thus, a taste reactivity test was conducted<br />

in a separate group of CTA rats. ID 0.3M sucrose preload<br />

(n=6) significantly increased gaping to intraoral 0.3M sucrose<br />

infusions (0.5ml/30s every 3min <strong>for</strong> 15min) relative to ID 0.15M<br />

NaCl preload (n=7). Together, the results suggest postoral<br />

stimuli impact oral taste processing in chemospecific ways.<br />

Acknowledgements: NIH-T32DC000044<br />

#P177 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

An interaction of chronic and acute metabolic states on<br />

olfactory perception associated with ghrelin signaling<br />

Xue Sun, Marga G Veldhuizen, Dana M Small<br />

The John B. Pierce Laboratory and Yale University New Haven,<br />

CT, USA<br />

Olfaction is key <strong>for</strong> the evaluation of food cues in the<br />

environment. It is there<strong>for</strong>e not surprising that evidence is<br />

emerging <strong>for</strong> metabolic influences in olfaction. For example,<br />

receptors <strong>for</strong> feeding-related molecules such as ghrelin are<br />

expressed within the olfactory system and olfactory sensitivity<br />

is increased in fasted vs. fed state in rodents. However, human<br />

studies have generated conflicting results. This study tested<br />

whether the influence of acute homeostatic state (fasted vs.<br />

fed) on olfactory perception could be moderated by body mass<br />

index, which is associated with chronic metabolic alterations.<br />

Twenty-one subjects (14 healthy weight/HW, BMI M=22.3<br />

SD=1.9; 7 overweight/OW, BMI M=28.2 SD=2.8) rated the<br />

intensity of odors (chocolate, strawberry, honeysuckle and<br />

lilac) and tastes (chocolate and strawberry flavored milk) using<br />

the general labeled magnitude scale in fed and fasted states<br />

(separate counterbalanced days), concomitant to an fMRI<br />

study. Blood samples were also collected on a subset of subjects<br />

(n=15) to assess the influence of internal state on feeding-related<br />

molecules. A repeated measures ANOVA of intensity<br />

ratings revealed a 3-way interaction between state<br />

(fed vs. fasted), stimulus (odors vs. taste) and group (HW vs.<br />

OW) (f(1,19)=7.52, p=.013), where odor intensity ratings<br />

decreased significantly in fed vs. fasted <strong>for</strong> OW but not HW<br />

subjects. Critically, the difference in odor intensity ratings<br />

between fasted and fed correlated negatively with difference in<br />

ghrelin change from baseline between fasted and fed conditions<br />

(r(13)=-.615, p=.016); the smaller the change in intensity<br />

perception, the larger the change in ghrelin levels. These findings<br />

demonstrate that chronic and acute metabolic states interact to<br />

influence olfactory perception possibly through ghrelin signaling.<br />

Acknowledgements: Supported by R01 DK085579.<br />

#P178 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

Oral Sweet Taste Stimulation Induces Cephalic Phase<br />

Carbohydrate Oxidation in Humans<br />

Sze Yen Tan, Richard D. Mattes<br />

Purdue University/Department of Nutrition Science West<br />

Lafayette, IN, USA<br />

Objective: Sensory stimuli induce many anticipatory reflexes.<br />

Specifically, oral sweet taste stimulation increases glucose<br />

absorption in rats and induces cephalic phase insulin release<br />

in animals and humans. This study aimed to investigate<br />

whether oral sweet taste stimulation also induces an acute<br />

increase of carbohydrate oxidation in humans. Methods: This<br />

was a randomized, crossover design study with two test visits<br />

delivering: 1) control (DI water) and 2) 10% sucrose (w/w)<br />

solutions orally. An indirect hood calorimeter was used to<br />

measure energy expenditure (EE) and respiratory quotient (RQ)<br />

through gaseous exchanges. Each measurement was per<strong>for</strong>med<br />

at one to two hours after habitual breakfast or lunch, where<br />

participants were instructed to consume an identical meal at<br />

identical times be<strong>for</strong>e their visits. Calorimeter measurement<br />

was preceded by a 30-minute habituation period, followed by<br />

oral stimulation and subsequent measurement <strong>for</strong> 30 minutes.<br />

Changes in EE and RQ were tested using a general linear<br />

model <strong>for</strong> repeated measures ANOVA in SPSS. Results: Nine<br />

participants have completed the study (8 females and 1 male,<br />

mean age=25.2 ± 4.6 years, mean weight=59.2 ± 4.3kg, mean<br />

BMI=21.6 ± 2.5kgm -2 ). Energy expenditure following test<br />

solution stimulation did not change and was not significantly<br />

different between solutions (time and interaction effects, P>0.05).<br />

However, RQ was significantly increased after sweet taste<br />

stimulation (time effects, p


#P179 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

#P180 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

Energy supply in chemosensory cilia of olfactory receptor<br />

neurons: Possible role of glycolysis<br />

Pablo S Villar 1,2 , Juan G Reyes 1 , Juan Bacigalupo 2<br />

1<br />

Instituto de Química, Facultad de Ciencias, Pontificia Universidad<br />

Católica de Valparaíso Valparaíso, Chile, 2 Departamento de Biología,<br />

Facultad de Ciencias, Universidad de Chile Santiago, Chile<br />

The chemosensory cilia of olfactory sensory neurons are long<br />

and thin structures ( 60 x 0.2 µm) devoid of inner membranes,<br />

specialized in odorant transduction. A cAMP pathway couples<br />

the activation of odor receptors with the opening of cyclic<br />

nucleotide-gated channels. During the odor response, the cilia<br />

undergo high levels of ATP hydrolysis, as this nucleotide is<br />

used by adenylyl cyclase, ATPases and kinases. Our estimates<br />

of resting ATP level, ATP diffusion and consumption suggest<br />

that the mitochondria, located near the base of the cilia, are<br />

insufficient to sustain chemotransduction in the entire cilium<br />

under intensive stimulation. Nuñez-Parra et al (Chem Senses<br />

36:771-7802012) found glucose transporters in the sustentacular<br />

cells of the olfactory epithelium. We hypothesize these cells<br />

release glucose to the mucus, the cilia incorporated it from there<br />

and utilize it by glycolysis, supplementing the required ATP.<br />

To test this idea, we detected glycolytic enzymes by immunoblot<br />

of a ciliary membrane preparation. Additionally, we measured<br />

cilia and knob accumulation of a fluorescent deoxyglucose<br />

analog when applied to mucosal side of the olfactory epithelium,<br />

suggesting the apical presence of a glucose transporter. We<br />

demonstrated by immunocytochemistry the ciliary location<br />

of this transporter in isolated rat and toad olfactory neurons.<br />

Additionally, field recordings (electroolfactogram) indicated that<br />

inhibition of glycolysis and oxidative phosphorylation impairs<br />

the odor response. Altogether, these results are consistent with a<br />

dual supply of ATP in olfactory cilia, oxidative phosphorylation<br />

and glycolysis. Acknowledgements: FONDECYT 1100682 (JB),<br />

DI/VRIEA/PUCV (JGR)<br />

Dietary calcium intake and ethnicity may contribute to<br />

individual differences in taste perception.<br />

Anna Voznesenskaya, Laura K. Alarcón, Michael G. Tordoff<br />

Monell Chemical Senses Center Philadelphia, PA, USA<br />

Calcium status affects preferences <strong>for</strong> calcium and sweet<br />

solutions in rats and mice: calcium deficient rodents drink<br />

more calcium solutions and avoid sweet compounds. Moreover,<br />

preference <strong>for</strong> calcium is inversely correlated with preference<br />

<strong>for</strong> sweet compounds in calcium replete mice. However, little is<br />

known about the relationships between calcium status and taste<br />

perception in humans. Here we measured detection thresholds<br />

<strong>for</strong> CaCl 2<br />

and sucrose, and assessed intensity and taste quality<br />

ratings of CaCl 2<br />

, sucrose, NaCl, QHCl and citric acid in an<br />

ethnically diverse group of people in relation to dietary calcium<br />

intake. African-Americans had significantly higher detection<br />

thresholds than Caucasians <strong>for</strong> both CaCl 2<br />

and sucrose. They<br />

rated 25 mM CaCl 2<br />

as predominantly sour significantly more<br />

frequently than Caucasians. There was no relationship between<br />

dietary calcium intake and CaCl 2<br />

detection threshold. In African-<br />

Americans but not Caucasians sucrose detection threshold was<br />

inversely correlated with dietary calcium levels (Spearman’s<br />

rho = - 0.93, p


#P181 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

#P182 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

Differences in food cue reactivity between normal weight<br />

and overweight individuals?<br />

Harriët F.A. Zoon 1 , Wei He 1,2 , Rene A. de Wijk 2 , Cees de Graaf 1 ,<br />

Sanne Boesveldt 1<br />

1<br />

Division of Human Nutrition, Wageningen University Wageningen,<br />

Netherlands, 2 Food and Biobased Research, Wageningen UR<br />

Wageningen, Netherlands<br />

Overweight occurs when energy intake exceeds energy<br />

expenditure over the long term. Overweight people have been<br />

suggested to be more sensitive to rewarding effects of food (e.g.<br />

Davis et al., 2004; Franken & Muris, 2005). In the anticipatory<br />

phase of eating, odors can be considered as external cues that<br />

signal the energy content and hence the reward value of a food.<br />

In overweight individuals internal hunger signals are thought<br />

to be overruled by external food cues (Herman & Polivy, 2008).<br />

This study aims to determine if food cue reactivity is higher in<br />

overweight compared to normal weight individuals. Frequency<br />

of choice <strong>for</strong> energy-dense food items and amount of food intake<br />

reflect food cue reactivity. 25 overweight (BMI mean: 31.33 kg/<br />

m2, SD: 3.36) and 25 normal weight (BMI mean: 21.84 kg/<br />

m2, SD: 1.78) females, matched on age and restraint score,<br />

participated. In 6 separate sessions they were exposed to odors<br />

of three different categories (signaling non-food, high-energy<br />

food, low-energy food) in two motivational states (hungry<br />

and satiated). High-energy preference was measured with a<br />

computerized <strong>for</strong>ced choice task and food intake (kCal) was<br />

determined with the use of a Bogus Taste Test. We hypothesize<br />

that increased food cue reactivity in overweight women is<br />

demonstrated by a stronger tendency to choose high-energy food<br />

products after being exposed to high-energy food odors, and<br />

may subsequently lead to more food intake compared to lean<br />

individuals. However, preliminary results (N=28) indicate that<br />

there is no main effect of odor on high-energy food preference<br />

(p=0.755) and also no interaction effect between odor and BMI<br />

group (p=0.935). Our first results on food intake (N=48) indicate<br />

no main effect of odor (p= 0.792) and no interaction effect<br />

of odor and BMI group (p=0.323). Acknowledgements: This<br />

study was funded by NWO (The Netherlands Organization <strong>for</strong><br />

Scientific Research), Veni grant nr. 451-11-021, awarded to SB.<br />

Development of Viral Based Gene Delivery <strong>for</strong> Conditional<br />

Ablation of Specific Brain Peptidergic Neurons<br />

Ali Magableh, Robert Lundy<br />

University of Louisville School of Medicine/Anatomical <strong>Sciences</strong> and<br />

Neurobiology Louisville, KY, USA<br />

Learning plays a crucial role in the establishment and<br />

strengthening of food preference and we hypothesize that<br />

specific limbic system neuropeptide pathways play an important<br />

role. Identifying neural mechanisms that mediate affective<br />

aspects of taste perception will further our understanding<br />

of how the brain controls eating and overeating. We have<br />

identified two neuropeptides, corticotrophin releasing factor<br />

(CRF) and somatostatin (Sst), which are expressed in limbic<br />

system neurons that project to a hindbrain neural substrate<br />

critical <strong>for</strong> establishment of gustatory hedonic value; the pontine<br />

parabrachial nucleus. Our goal is to develop a viral construct<br />

capable of directing conditional expression of nitroreductase<br />

gene (NTR) to Sst and CRF cell populations in the limbic<br />

system of mice using a cre/lox system. Thus, specific peptide<br />

producing neurons can be rapidly ablated in isolation following<br />

treatment with the prodrug CB1954 allowing assessment of<br />

their role in central taste processing and taste-guided behaviors.<br />

In vitro cell culture of HEK293 cells combined with FLOW<br />

cytometric analysis indicate that we can conditionally express<br />

NTR and cause cell death following CB1954 treatment.<br />

Acknowledgements: This research work was supported by a<br />

grant from the Kentucky Science and Engineering Foundation as<br />

per Grant Agreement #KSEF-148-502-11-277 with the Kentucky<br />

Science and Technology Corporation.<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

98


#P183 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

#P184 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

Combinatorial and genotype specific co-expression of the<br />

major urinary proteins (MUPs) during mouse postnatal<br />

development: from fundamental aspects of olfaction to<br />

innovative prospects in biomedicine<br />

Sergey N. Novikov 1 , Elena M. Fedorova 1,2 , Irina I. Ermakova 3 , Anatoly<br />

A. Philimonenko 4 , Gennady A. Churakov 5 , Sergey V. Mylnikov 6<br />

1<br />

I.P. Pavlov Institute of Physiology, Russian Academy of <strong>Sciences</strong><br />

Saint Petersburg, Russia, 2 Institute of Experimental Medicine, Russian<br />

Academy of Medical <strong>Sciences</strong> Saint Petersburg, Russia, 3 Institute of<br />

Cytology, Russian Academy of <strong>Sciences</strong> Saint Petersburg, Russia,<br />

4<br />

Institute of Molecular Genetics, v.v.i., Academy of <strong>Sciences</strong> of the<br />

Czech Republic Prague, Czech Republic, 5 Institute of Experimental<br />

Pathology/Molecular Neurobiology (ZMBE), University of Muenster<br />

Muenster, Germany, 6 Saint Petersburg State University, Department of<br />

Genetics and Biotechnology Saint Petersburg, Russia<br />

Major urinary proteins (MUPs) of the house mouse <strong>for</strong>m a large<br />

group of highly polymorphic acidic iso<strong>for</strong>ms with molecular<br />

masses of 18-20 kDa. MUPs are encoded by the Mup gene<br />

cluster, which consists of about 35 genes and pseudogenes and<br />

is mapped to chromosome 4. Nowadays MUPs are considered<br />

as a key component of the mouse olfactory signature which<br />

can provide all essential in<strong>for</strong>mation about the individuality of<br />

donors. There is also rapidly growing evidence that several MUP<br />

iso<strong>for</strong>ms are involved in the regulation of glucose metabolism<br />

(Zhou, Rui, 2010) and can be used as sensitive biomarkers in<br />

early diagnosis of experimental nephritis (Wenderfer et al., 2009)<br />

and hepatocarcinogenesis (Ritorto, Borlak, 2011). These studies<br />

open practically unexplored biomedicine avenue <strong>for</strong> using MUPs<br />

as new protein markers which are very suitable <strong>for</strong> diagnostic<br />

purposes. We examined ontogenetic profiles of MUPs expression<br />

in male and female mice of CBA/LacY and C57BL/6JY<br />

strains using electrophoresis in polyacrylamide gel (PAGE).<br />

Quantitative evaluation of eight MUPs iso<strong>for</strong>ms (A-H) revealed<br />

that each genotype is characterized by specific combinations<br />

and different proportions (ratios) of the same MUP fractions.<br />

These sex and genotype specific ratios emerged in both sexes<br />

very soon after weaning, remain quite constant in adults and<br />

resemble «barcode». Our data suggest that the pattern of Mup<br />

genes expression during mouse ontogenesis is regulated through<br />

a very stable genetic program. We suppose that at the early<br />

stage of illness this ontogenetic program is destroyed and the<br />

expression pattern of several Mup genes will be changed. These<br />

processes are reflected in the appearance of new protein profiles<br />

with altered MUPs’ ratios and may correspond to epigenetically<br />

changed expression of the Mup gene cluster. Acknowledgements:<br />

Supported by Russian Foundation <strong>for</strong> Basic Research (projects<br />

02-04-49273, 07-04-01762).<br />

The <strong>Association</strong> of Taste with Adiposity in the Beaver<br />

Dam Offspring Study<br />

Mary E. Fischer 1 , Karen J. Cruickshanks 1,2 , Carla R. Schubert 1 ,<br />

Guan-Hua Huang 3 , Barbara E.K. Klein 1 , Ronald Klein 1 ,<br />

James S. Pankow 4 , Nathan Pankratz 5 , Alex Pinto 1<br />

1<br />

University of Wisconsin, Department of Ophthalmology & Visual<br />

<strong>Sciences</strong> Madison, WI, USA, 2 University of Wisconsin, Department<br />

of Population Health <strong>Sciences</strong> Madison, WI, USA, 3 National Chiao<br />

Tung University, Institute of Statistics Hsinchu, Taiwan, 4 University<br />

of Minnesota, Division of Epidemiology & Community Health<br />

Minneapolis, MN, USA, 5 University of Minnesota, Department of<br />

Laboratory Medicine & Pathology Minneapolis, MN, USA<br />

Taste sensation may influence food choice and consumption<br />

which in turn may play a role in the maintenance of health.<br />

The objective of this study was to determine the relationship<br />

between taste and changes in adiposity and related health<br />

measures during a 5 year follow-up period in the Beaver Dam<br />

Offspring Study. Whole mouth suprathreshold taste intensity was<br />

measured <strong>for</strong> salt, sweet, sour, and bitter at baseline (2005-2008)<br />

using filter paper disks and a general labeled magnitude scale.<br />

Health outcomes measured at baseline and follow-up (2010-<br />

2013) included body mass index (BMI), waist circumference,<br />

systolic and diastolic blood pressure, total cholesterol, and<br />

hemoglobin A1c (HbA1c). Cluster analysis was used to group<br />

participants according to observed patterns of intensities of the 4<br />

tastes. In preliminary analyses (n = 1681, mean age at baseline =<br />

48.9 years, range = 22-84 years), there were associations between<br />

patterns of taste intensities and 5-year changes in BMI, waist<br />

circumference, and HbA1c level. With adjustment <strong>for</strong> age and<br />

sex, the cluster with high intensities <strong>for</strong> all 4 tastes demonstrated<br />

a significantly greater mean increase in BMI (+ 0.96 kg/m 2 )<br />

and HbA1C (+ 0.37%) than the cluster with average intensities<br />

<strong>for</strong> the 4 tastes (BMI: + 0.32 kg/m 2 ; HbA1c: + 0.21%). Similar<br />

results were observed <strong>for</strong> waist circumference (high intensities<br />

cluster: + 3.01 cm; average intensities cluster: + 1.87 cm). In<br />

these preliminary analyses, oral sensation, characterized using<br />

patterns of perceived intensities of suprathreshold tastes, was<br />

found to be associated with 5-year changes in some adiposityrelated<br />

health outcomes. Acknowledgements: The project<br />

described was supported by R01AG021917 from the National<br />

Institute on Aging, National Eye Institute, and National Institute<br />

on Deafness and Other Communication Disorders. The content<br />

is solely the responsibility of the authors and does not necessarily<br />

reflect the official views of the National Institute on Aging or the<br />

National Institutes of Health.<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

99


#P185 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

#P186 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

Social Olfactory Cues and Stress<br />

Pamela Dalton, Cristina Jaen, Tamika Wilson, Christopher Maute<br />

Monell Chemical Senses Center Philadelphia, PA, USA<br />

Olfactory cues have the potential to precipitate emotional<br />

responses and thereby alter mood, judgments and behavior.<br />

In prior work, we demonstrated that exposure to a novel odor<br />

while undergoing a laboratory stressor caused individuals<br />

to re-experience stress (increased heart rate & self-reported<br />

stress) when re-exposed to that odor three days later. We<br />

wished to evaluate the potential <strong>for</strong> odor to alter stress levels<br />

following a standard laboratory stressor. One class of olfactory<br />

stimuli which has shown promise in eliciting robust effects on<br />

mood and emotion are social odors. Lundström et al. (2008)<br />

demonstrated that smelling a stranger’s body odor activated<br />

a marked response in the amygdala of subjects, despite a low<br />

conscious recognition of the odor or its source. In this study,<br />

we evaluated the changes in autonomic stress levels following<br />

the Trier Social Stress Test among individuals in 3 groups who<br />

were exposed to either the body odor of their sibling, the body<br />

odor of a stranger or a non-social (fragrance) odor. Axillary<br />

odors were collected from non-twin, whole, biological siblings<br />

who were then recalled to participate in the main study in which<br />

one of the 3 odors was administered following the stress task.<br />

Results showed a significant decrease in post-recovery heart<br />

rate only among the group smelling the sibling odor, whereas<br />

skin conductance was significantly reduced <strong>for</strong> both the sibling<br />

odor and the fragrance. Following a stressor, exposure to the<br />

stranger body odor maintained arousal levels longer suggesting<br />

that both familiar and stranger body odors may be potent cues<br />

<strong>for</strong> emotional responses. Acknowledgements: Supported by the<br />

U.S. Army Research Office grant # W911NF-11-1-0087, entitled<br />

“Learning & Olfaction: Understanding and Enhancing a Critical<br />

Communication Channel”.<br />

Can a chemosensory threat be masked?<br />

Amy R Gordon 1,2 , Kathrin Ohla 2,3 , Mats J Olsson 1 ,<br />

Johan N Lundstrom 1,2,4<br />

1<br />

Karolinska Institutet Stockholm, Sweden, 2 Monell Chemical Senses<br />

Center Philadelphia, PA, USA, 3 German Institute of Human<br />

Nutrition Potsdam-Rehbrücke, Germany, 4 University of<br />

Pennsylvania Philadelphia, PA, USA<br />

The well-established angry advantage effect has been extended<br />

in recent crossmodal visual-olfactory studies using schematic<br />

human faces and human body odors. In a threat-detection task,<br />

humans detect an angry (threatening) schematic face in an<br />

array of neutral distracter faces more quickly than a friendly<br />

(non-threatening) face, hence the label ‘Angry advantage’.<br />

We have previously shown that the body odor of unknown<br />

individuals (Strangers) – an established threatening olfactory<br />

stimulus – speeds a subject’s detection of threatening faces, but<br />

not non-threatening faces, relative to exposure to the subject’s<br />

own body odor (Self). Using event-related potential (ERPs),<br />

we have more recently demonstrated that the presence of<br />

a Strangers’ body odor causes a non-threatening face to be<br />

processed as a threatening stimulus. In the present ERP study,<br />

we sought to determine whether the chemosignal mediating the<br />

a<strong>for</strong>ementioned ERP effect could be masked by a common odor<br />

(Mask). Angry and neutral schematic faces were presented to<br />

subjects in the presence of Strangers’ body odor + Mask, ‘Self’<br />

body odor + Mask, or Mask only control, which were delivered<br />

intra-nasally by a computer-controlled olfactometer. Preliminary<br />

analyses suggest that even in the presence of an odor mask,<br />

exposure to the body odor of a stranger, relative to the odorless<br />

control and ‘Self’ body odor, results in significant differences<br />

in the late (cognitive) components of visual processing. This<br />

suggests that body odor can modulate the cognitive evaluation<br />

of visual stimuli even in the presence of a perceptual odor mask.<br />

The effects of masked body odor and common odor exposure<br />

on visual processing will be presented and discussed within the<br />

framework of the adaptive advantages conveyed by heightened<br />

sensitivity to threat-related stimuli. Acknowledgements: This<br />

work was supported by the National Institute on Deafness and<br />

other Communication Disorders – NIDCD (R03DC009869)<br />

awarded to JNL<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

100


#P187 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

Differential responses to two kairomonal cues in mosquitoes<br />

Shahid Majeed, Sharon R. Hill, Teun Dekker, Göran Birgersson,<br />

Rickard Ignell<br />

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

#P188 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

Effects of 5a-Androst-16-EN-3a-OL Scent Administration<br />

on Augmenting Male Gambling Behavior<br />

Sierra Moore, Bryan Raudenbush, Patrick Dwyer<br />

Wheeling Jesuit University Wheeling, WV, USA<br />

Culex quinquefasciatus, Aedes aegypti, and Anopheles gambiae are<br />

vectors of diseases that are among the main causes of human<br />

mortality and morbidity worldwide. Host-seeking in these species<br />

is primarily regulated by olfactory cues. The most important cue<br />

<strong>for</strong> the activation of host-seeking is carbon dioxide (CO 2<br />

), the<br />

principal by-product of respiration. All three mosquito species<br />

were able to detect and follow pulsed stimuli of CO 2<br />

at the level<br />

of olfactory receptor neurons (ORNs) housed within capitate<br />

pegs on the maxillary palps. The temporal coding capacity of C.<br />

quinquefasciatus CO 2<br />

-sensitive ORNs, however, was significantly<br />

lower than that of the other two species. This differential<br />

physiological response was reflected in the behavioral response<br />

to CO 2<br />

, and correlates with the CO 2<br />

emissions from the preferred<br />

hosts <strong>for</strong> each of these species. Furthermore, aeration extracts<br />

taken from preferred hosts were analyzed by gas chromatography<br />

coupled single sensillum recording (GC-SSR) of the capitate<br />

pegs. We identified (R)-1-octen-3-ol, a component in human<br />

headspace volatiles, as a physiologically active in each species,<br />

although with different sensitivities. It is interesting to note that<br />

(R)-1-octen-3-ol was absent from bird aeration extracts. Landing<br />

bioassays using the host aeration extracts revealed behavioral<br />

responses of the three species consistent with their host selection<br />

preferences. The addition of biologically relevant concentrations<br />

of (R)-1- octen-3-ol to bird aeration extracts either inhibited or<br />

increased the behavioral response of the mosquitoes, consistent<br />

with its role as a non-host and host volatile, respectively. Here,<br />

we show that the host-seeking behavior of mosquitoes may be<br />

differentially regulated by olfactory signals emitted by potential<br />

hosts in their environment.<br />

The effects of pheromones on non-humans have been extensively<br />

studied. However, less is known on how such scents can elicit<br />

responses in humans. Previous research indicates that the<br />

scent of a female can affect various emotions and behaviors of<br />

males. The present study assessed the effects of Androstenol<br />

on gambling behavior in males. Participants completed an ad<br />

libitum blackjack gambling task while exposed to either no<br />

scent (control) or androstenol (experimental condition). Results<br />

indicate males gambled <strong>for</strong> a significantly longer period of time<br />

when exposed to androstenol, t(36) = 2.09, pA), which changes<br />

amino acid 180 in the resultant protein’s polypeptide chain from<br />

glycine (G) to arginine (R), were found to have significantly less<br />

of the characteristic axillary odorants than either those who<br />

are heterozygotic <strong>for</strong> this change or those who had the wild<br />

type gene. Asian populations differ markedly from non-Asians<br />

in their ear wax type and underarm odor production. G180R<br />

SNP also is associated with a dry, white earwax phenotype that<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

101


is predominant in Asians. The G180R SNP is rare in Africans<br />

and Caucasians who typically exhibit wet, yellow earwax. For<br />

the first time, analytical analysis of earwax odorants has been<br />

per<strong>for</strong>med and the principle odorants in both earwax phenotypes<br />

will be discussed. The odor of each ear wax type was in<strong>for</strong>mally<br />

accessed and the principal odorants were found to be volatile<br />

organic C 2<br />

-to-C 8<br />

acids. A comparison between volatile earwax<br />

and axillary odors will also be presented. Acknowledgements:<br />

NIH postdoctoral training grant (2T32DC000014-32A1) ARO<br />

(W911NF-11-1-0087)<br />

#P190 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

Loss and Recovery of Odorant-Mediated Behavior Correlates<br />

with Plasticity of Axonal Projections in the Zebrafish<br />

Olfactory Bulb in a Reversible Deafferentation Model<br />

Evan J White, Taylor R Paskin, Christine A Byrd-Jacobs<br />

Western Michigan University/Biological <strong>Sciences</strong> Kalamazoo, MI, USA<br />

We have found that the repeated exposure of adult zebrafish<br />

olfactory epithelium to the detergent Triton X-100 results in fish<br />

losing the ability to respond to odorants associated with social<br />

behavior but retaining the ability to respond to odorants linked<br />

to feeding behavior. Using a reversible deafferentation technique,<br />

we find that fish recover the ability to detect social cues. The<br />

aim of the present study was to determine a biological basis<br />

<strong>for</strong> this phenomenon by examining axonal projections after a<br />

single treatment with TX-100. Axons of three olfactory sensory<br />

neuron subtypes (ciliated, microvillar, and crypt) were identified<br />

using immunocytochemistry on paraffin sections. In control<br />

bulbs, anti-KLH labeled all glomeruli, while anti-calretinin<br />

labeled fewer axons throughout the bulb. Anti-Gas/olf labeling<br />

was concentrated in the medial and dorsal bulb, and anti-S-100<br />

labeling was more obvious in the lateral bulb. Within the first<br />

4 days after TX-100 treatment, anti-KLH and anti-calretinin<br />

labeling in the deafferented bulb showed an overall reduction,<br />

with prominent loss in the medial bulb and preservation of some<br />

axons in the lateral bulb. By 7 days, innervation returned to near<br />

control levels. Staining in the deafferented bulb with anti-Gas/olf<br />

and anti-S-100 was absent 1 day following treatment but returned<br />

within 7 days. Examination of the axon patterns showed a<br />

selective preservation of certain olfactory sensory axons, while<br />

others are temporarily destroyed. The presumptive microvillar<br />

axons that survive treatment in the lateral bulb may account <strong>for</strong><br />

the persistent ability of zebrafish to detect food odorants while<br />

the temporary destruction of ciliated axons in the medial bulb<br />

is consistent with the loss and recovery of the ability to detect<br />

social cues. Acknowledgements: Supported by NIH-NIDCD<br />

#011137 (CBJ)<br />

#P191 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

Coexistence of determined and variable sensory coding<br />

strategies in the mouse vomeronasal system<br />

Tobias Ackels 1 , Annika Cichy 1 , Angeldeep Kaur 2 , Maria Kateri 3 ,<br />

Tobias Marton 2 , Darren Logan 2,4 , Lisa Stowers 2 , Marc Spehr 1<br />

1<br />

Department of Chemosensation, RWTH Aachen University Aachen,<br />

Germany, 2 Department of Cell Biology, The Scripps Research Institute<br />

La Jolla, CA, USA, 3 Institute of Statistics, RWTH Aachen University<br />

Aachen, Germany, 4 Wellcome Trust Sanger Institute, Hinxton<br />

Cambridge, United Kingdom<br />

The mouse vomeronasal organ (VNO) is an important<br />

chemosensory subsystem that has been implicated in a variety<br />

of social and sexual behaviors. In contrast to combinatorial<br />

odor coding by neurons in the main olfactory epithelium,<br />

vomeronasal sensory neurons (VSNs) are thought to function<br />

as narrowly tuned, dedicated sensors of intrinsically instructive<br />

semiochemicals. Here, we investigate the tuning profile(s)<br />

of a group of VSNs that are collectively characterized by<br />

their sensitivity to a specific class of behaviorally relevant<br />

chemosignals: major urinary proteins (MUPs). Using<br />

extracellular ‘loose-seal’ patch-clamp recordings from optically<br />

identified basal VSNs in acute coronal VNO slices, we record<br />

stimulus-dependent action potential discharge in response to<br />

recombinant MUPs, specific <strong>for</strong> either the C57BL/6J or the<br />

BALB/cByJ inbred strain of laboratory mice. Furthermore,<br />

we comparatively analyzed the role(s) of these stimuli in two<br />

different male-specific behaviors: male-male aggression and<br />

territorial countermarking. Surprisingly, electrophysiological<br />

activity profiling revealed parallel detection of MUPs by both<br />

‘specialist’ neurons selectively tuned to a particular stimulus<br />

and broad range responders (‘generalists’) sensitive to all or<br />

subset combinations of the MUPs tested. These data suggest the<br />

coexistence of determined and variable sensory coding strategies<br />

in the mouse vomeronasal system. In addition, behavioral assays<br />

indicate that MUPs regulate at least two different male behaviors.<br />

While dedicated ligands promote aggression, a combinatorial<br />

MUP code controls countermarking. Together, our results show<br />

that a vomeronasal stimulus can encode divergent in<strong>for</strong>mation<br />

through both dedicated and combinatorial neural mechanisms.<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

102


#P192 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

#P193 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

Insights into the Function of Darcin from the Three<br />

Dimensional Structure<br />

Robert J Beynon, Marie M Phelan, Lu-Yun Lian, Lynn McLean,<br />

Jane L Hurst<br />

University of Liverpool / Institute of Integrative Biology Liverpool,<br />

United Kingdom<br />

Mouse urine contains millimolar concentrations of major urinary<br />

proteins (MUP) - eight stranded beta barrel lipocalins. MUPs<br />

have a broad range of functions in chemical communication<br />

between individuals. One MUP, darcin (MGI classification<br />

Mup20) is highly expressed in males only and is responsible <strong>for</strong><br />

inherent female attraction to males, the subsequent memory of<br />

the volatile odour profile of that male and the rapid induction<br />

of memory of the precise physical placement of the scent mark<br />

containing darcin. Darcin is responsible <strong>for</strong> the slow release<br />

of the volatile pheromone, 2-sec-butyl 4,5 dihydrothiazole. To<br />

better understand the unique properties of darcin, we have<br />

solved the structure of this protein by NMR. Relative to other<br />

MUPs, darcin has a large solvent exposed area and the greatest<br />

exposure of hydrophobic residues in the beta barrel. Binding<br />

to three ligands – NPN, menadione and a thiazole derivative<br />

– were characterized in this study. Menadione and thiazole<br />

bound to both darcin and MUP11 in the hydrophobic cavity<br />

of the beta barrel, with NMR data indicating a similar binding<br />

site <strong>for</strong> menadione and thiazole in both darcin and MUP11.<br />

The largest ligand (NPN) bound only to MUP11, not darcin,<br />

suggesting darcin adopts a more compact binding cavity. Darcin<br />

is significantly more stable than MUP11, with 92% of darcin<br />

structure retained in the native state at 7M urea compared to<br />

only 45% in MUP11. The high stability of darcin is consistent<br />

with the anomalous migration on SDS-PAGE and a tendency to<br />

undercharge in electrospray ionisation mass spectrometers. All<br />

this biophysical and ligand binding data point to darcin adopting<br />

a more stable and compact con<strong>for</strong>mation with a smaller ligand<br />

binding cavity than related MUPs. Acknowledgements: These<br />

studies were supported by the Biotechnology and Biological<br />

Science Research Council (BB/J002631/1).<br />

HCN Channels Mediate Proton-dependent Signaling in the<br />

Mouse Vomeronasal Organ<br />

Annika Cichy, Tobias Ackels, Jennifer Spehr, Marc Spehr<br />

RWTH Aachen University, Dept. Chemosensation Aachen, Germany<br />

The mouse vomeronasal organ (VNO) plays an important role in<br />

the detection of semiochemicals and other social cues. However,<br />

many of the basic mechanisms that control VNO physiology<br />

remain largely unknown. Here, we investigate proton-mediated<br />

activity in the mouse VNO. We show that mouse urine is not<br />

only a rich source of social chemosignals, but can also create<br />

an acidic environment <strong>for</strong> such cues. For females, in particular,<br />

we find an experience-dependent variation of their generally<br />

low urinary pH. Using whole-cell patch-clamp recordings from<br />

visually identified sensory neurons in acute tissue slices of<br />

the mouse VNO, we show that vomeronasal sensory neurons<br />

are activated by protons. We describe that acidic solutions<br />

dose-dependently induce inward currents in voltage-clamp<br />

measurements and elicit robust action potential firing in currentclamp<br />

recordings. Surprisingly, our investigations suggest<br />

no substantial involvement of ‘classical’ candidate protonactivated<br />

ion channels. Instead, the pharmacological profile and<br />

biophysical properties of the proton-induced responses indicate<br />

a critical role of hyperpolarization-activated cyclic-nucleotidegated<br />

(HCN) ion channels in proton-mediated signaling of<br />

vomeronasal sensory neurons. Together, our results implicate<br />

HCN channel-dependent vomeronasal acid-sensing in gain<br />

control of social chemosignaling.<br />

#P194 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

Kirrel-3 is Required <strong>for</strong> the Coalescence of Vomeronasal<br />

Sensory Neuron Axons into Glomeruli and <strong>for</strong> Male-Male<br />

Aggression<br />

Jean-François Cloutier 1,2 , Alexandra Brignall 1,2 , Tyler Cut<strong>for</strong>th 3 ,<br />

Kang Shen 4 , Janet Prince 1,2<br />

1<br />

Montreal Neurological Institute Montréal, QC, Canada, 2 McGill<br />

University Montréal, QC, Canada, 3 University of Cali<strong>for</strong>nia Irvine,<br />

CA, USA, 4 Stan<strong>for</strong>d University Stan<strong>for</strong>d, CA, USA<br />

POSTER PRESENTATIONS<br />

The accessory olfactory system controls social and sexual<br />

interactions in mice that are critical <strong>for</strong> survival. Vomeronasal<br />

sensory neurons (VSNs) <strong>for</strong>m synapses with dendrites of<br />

second order neurons in glomeruli of the accessory olfactory<br />

bulb (AOB). Axons of VSNs expressing the same vomeronasal<br />

receptor (VR) coalesce into multiple glomeruli within spatially<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

103


conserved regions of the AOB. Here we examine the role of the<br />

Kirrel family of transmembrane proteins in the coalescence of<br />

VSN axons within the AOB. We find that Kirrel-2 and Kirrel-3<br />

are differentially expressed in subpopulations of VSNs and<br />

that their expression is regulated by activity. While Kirrel-3<br />

expression is not required <strong>for</strong> early axonal guidance events,<br />

such as fasciculation of the vomeronasal tract and segregation<br />

of apical and basal VSN axons in the AOB, it is necessary <strong>for</strong><br />

proper coalescence of axons into glomeruli. Ablation of Kirrel-3<br />

expression results in disorganization of the glomerular layer of<br />

the posterior AOB and <strong>for</strong>mation of fewer, larger glomeruli.<br />

Furthermore, altered glomerular <strong>for</strong>mation is associated with<br />

loss of male-male aggression in kirrel-3 -/- mice. Taken together<br />

our results indicate that differential expression of Kirrels on<br />

vomeronasal axons generates a molecular code that dictates<br />

their proper coalescence into glomeruli within the AOB.<br />

Acknowledgements: Canadian Institutes of Health Research and<br />

Natural <strong>Sciences</strong> and Engineering Research Council of Canada.<br />

#P195 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

Transduction <strong>for</strong> pheromones in the main olfactory epithelium<br />

is mediated by the Ca 2+ - activated channel TRPM5<br />

Diego Restrepo 1 , Fabian Lopez 2 , Roberto Lopez 1 , Juan Bacigalupo 2<br />

1<br />

Department of Cell and Developmental Biology, Neuroscience Program<br />

and Rocky Mountain Taste and Smell Center, University of Colorado<br />

Anschutz Medical Campus Aurora, CO, USA, 2 Department of Biology,<br />

University of Chile Santiago, Chile<br />

The main olfactory epithelium contains olfactory sensory<br />

neurons (OSNs) that respond to odorants. Interestingly, there is<br />

growing evidence that some OSNs in this epithelium respond to<br />

pheromones through an unknown transduction mechanism. Here<br />

we report on a survey <strong>for</strong> pheromone transduction in a subset<br />

of OSNs expressing the transient receptor potential channel M5<br />

(TRPM5), a Ca 2+ -activated monovalent cation-selective channel.<br />

As in the majority of OSNs, the cyclic nucleotide-gated (CNG)<br />

channel subunit A2 is expressed in the cilia of OSNs expressing<br />

GFP under control of the TRPM5 promoter. Interestingly these<br />

TRPM5-GFP + OSNs lack the Ca 2+ -activated Cl - channel ANO2<br />

found in the majority of OSNs. Complementary loose patch<br />

current and Ca 2+ fluorescence recordings show that TRPM5-<br />

GFP + OSNs respond to pheromones and not to odorants, while<br />

TRPM5-GFP - OSNs respond to both. Finally complementary<br />

pharmacological and TRPM5 knockout experiments show<br />

that TRPM5-GFP OSNs respond to pheromones through the<br />

TRPM5 channel. Thus, pheromone responses of TRPM5-GFP +<br />

OSNs are mediated by ciliary Ca 2+ influx through CNG that<br />

gates opening of TRPM5. Acknowledgements: Funded by NIH<br />

DC006070 and DC004657 (D.R.), CONDECYT 1100682 (F.L.)<br />

and CONICYT and MECESUP UCH0713 (J.B.)<br />

#P196 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

AMYGDALAR PROCESSING OF SALIENT<br />

CHEMOSENSORY SIGNALS<br />

Lindsey M Biggs, Ariel R Simonton, Michael Meredith<br />

Florida State University/Program in Neuroscience, Dept. Biol.<br />

Sci. Tallahassee, FL, USA<br />

Medial amygdala responds differentially to conspecific and<br />

heterospecific chemosensory signals with different meanings<br />

and different behavioral responses; and it may be responsible <strong>for</strong><br />

routing in<strong>for</strong>mation to hypothalamic/preoptic circuits involved<br />

in producing the appropriate responses. The vomeronasal organ<br />

(VNO) is the primary but not only source of chemosensory<br />

input to medial amygdala but VNO-lesions disrupt the<br />

characteristic patterns of responses. The circuit <strong>for</strong> processing<br />

VNO-driven chemosensory input includes the main intercalated<br />

nucleus (mICN), one of several GABA-ir ICN cell groups<br />

in the amygdala. mICN appears to regulate posterior medial<br />

amygdala (MeP) activity similarly to the regulation of central<br />

and basolateral amygdala activity by paracapsular ICN cell<br />

groups, in the fear conditioning circuit. Using immediate-early<br />

gene expression, we previously found that GABA-receptor-ir<br />

cells in MeP are suppressed by heterospecific stimuli –as mICN<br />

GABA-ir cells are activated. Now using brain slice recording<br />

we show hyperpolarization of MeP cells by field stimulation<br />

of local mICN. Preliminary evidence also suggests mICN<br />

cells are suppressed by bath-applied dopamine, as is the case<br />

<strong>for</strong> paracapsular ICN cells. The amygdala contributes to the<br />

motivational/emotional evaluation of sensory inputs of all<br />

modalities and <strong>for</strong> diverse behaviors. This concordance between<br />

amygdala processing in completely different types of behavior<br />

may indicate some commonalities in circuit organization.<br />

Dopamine may be part of a mechanism <strong>for</strong> modulating<br />

amygdala processing of sensory in<strong>for</strong>mation according to brain<br />

state or previous experience. Dopamine appears to modulate<br />

experience-dependent chemosensory responses in basolateral<br />

amygdala but so far we have not demonstrated an effect in<br />

medial amygdala. Acknowledgements: Supported by NIDCD<br />

grants R01-DC005813, T32-DC000044 and funding from Florida<br />

State University<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

104


#P197 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

#P198 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

Interspecies communication mediated by tear fluids<br />

Mai Tsunoda, Kazushige Touhara<br />

Department of Applied Biological chemistry, The University of Tokyo<br />

Tokyo, Japan<br />

Communication between animals are regulated by a variety<br />

of chemical cues emitted from the body fluids. Recent works<br />

have revealed that exocrine grand-secreting peptide 1 (ESP1),<br />

which is released from male mouse tear fluids, enhances female<br />

sexual behavior through the vomeronasal organ. This data<br />

indicates that the tear fluid is one of the important sources of<br />

chemical cues. However, it is unknown whether tear-derived<br />

chemical cues mediate only intraspecies communication. In this<br />

study, we aimed to understand a novel function of tear fluids in<br />

interspecies communication by focusing on tear fluids of rats, a<br />

predator of mice. First, we examined the effect of rat tear fluids<br />

on the mouse vomeronasal system. c-Fos analysis revealed that<br />

rat tear fluids contained some stimulants that induced c-Fos<br />

expression in the accessory olfactory bulb (AOB), the first center<br />

of the vomeronasal system. It has been known that rats have 10<br />

members of ESP family genes, there<strong>for</strong>e, we examined whether<br />

the stimulants in rat tear fluids are ratESPs. Western blot analysis<br />

indicated that ratESP5 and ratESP7 were secreted in rat tear<br />

fluids. However, recombinant ratESP5 and ratESP7 did not<br />

induce c-Fos expression in the mouse AOB. This data suggests<br />

that there exists novel mouse vomeronasal stimulants in rat tear<br />

fluids. We next purified the stimulants from rat tear fluids by<br />

activity-based fractionation. Amino-terminal peptide sequence<br />

and genome analysis revealed that a c-Fos-inducing peptide<br />

was encoded by a gene whose function has not been revealed.<br />

We named this peptide P18. Recombinent P18 induced c-Fos<br />

expression in the AOB of wild type mice, but not in the TRPC2<br />

knock-out mice. These results suggest a possibility that P18 in<br />

rat tear fluids mediate interspecies communication through the<br />

vomeronasal organ.<br />

Experience-dependent plasticity causes sexual dimorphism<br />

in mouse pheromone-sensing neurons<br />

Pei S. Xu, Timothy E. Holy<br />

Department of Anatomy & Neurobiology, Washington University<br />

School of Medicine St. Louis, MO, USA<br />

In mice, the normal expression of most sex-specific behaviors<br />

requires an intact accessary olfactory system (AOS). While<br />

the AOS has been long viewed as a sexually dimorphic circuit,<br />

the known anatomical differences between males and females<br />

consist of modest changes in the packing of neurons in<br />

particular brain regions. By themselves, these differences may<br />

be insufficient to explain observed dimorphic behaviors. Here<br />

we asked whether the first order neurons, AOS sensory neurons<br />

differed functionally between two sexes. Using light-sheet based<br />

high-speed calcium imaging technique, we recorded ~260,000<br />

individual neurons in intact vomeronasal epithelia from male<br />

and female mice. According to the cell responses to 12 sulfated<br />

steroids, a class of chemicals that originally isolated from mouse<br />

urine, we classified a total of 20, 853 responsive neurons into 17<br />

functional types. We found that the large majority of functional<br />

receptor types present in equal abundance in males and females.<br />

However, we found clear sexual dimorphism, as two functional<br />

types appeared to be male specific, including an epitestosteroneselective<br />

receptor type 100-fold more abundant in males than in<br />

females. To explore the mechanism generating this dimorphism,<br />

we found male specific receptor types became rare after longterm<br />

exposure to the odors of female mice, with the result that<br />

the vomeronasal organs from males were converted to a pattern<br />

indistinguishable from females. This difference in AOS receptor<br />

type is by far the strongest sexual dimorphism ever reported in<br />

the mammalian central nervous system; that this dimorphism is<br />

determined entirely by experience indicates that a sensory system<br />

devoted to “innate” responses is strongly modulated by rearing<br />

conditions. Acknowledgements: This study was funded by NIH-<br />

NINDS/NIAAA Grant R01 NS068409, and NIH Director’s<br />

Pioneer Award DP1 OD006437(T.E.H.)<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

105


#P199 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

#P200 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

Palatability of Cycloheximide or Caffeine Mixed with<br />

Sugar in Hamsters<br />

Elizabeth M. Casey, Bradley K. Formaker, Thomas P. Hettinger,<br />

Marion E. Frank<br />

University of CT Health Center Farmington, CT, USA<br />

Multiple coding mechanisms <strong>for</strong> bitter stimuli are suggested<br />

by taste receptive-field specificity, quality and palatability.<br />

Quality and palatability of quinine (Qui), salicin (Sal), caffeine<br />

(Caf) and cycloheximide (Cyc) differ in hamsters (Mesocricetus<br />

auratus). Quality, studied with generalizations of conditioned<br />

taste aversions (CTA), differs <strong>for</strong> each; but, each stimulus is<br />

aversive measured in preference to water. Hamster taste aversions<br />

are generally reduced by adding sucrose. Thus, aversions<br />

to 1mM Qui, 10mM Sal, 30mM Caf and 30µM Cyc were<br />

compared to the 4 stimuli mixed with 500mM sucrose. Qui<br />

and Sal palatability increased with sucrose added but Caf and<br />

Cyc did not (Lloyd et al. 2012). To determine concentrationdependence,<br />

5, 10 and 30mM Caf and 5, 10 and 30mM Cyc<br />

(with and without 500mM sucrose) were tested <strong>for</strong> 2-bottle<br />

48-hr preference vs. water (R-L bottle positions reversed daily).<br />

Each hamster randomly received 14 stimuli based on a modified<br />

Latin Square. A preference ratio [ml stimulus ingested/ml total<br />

fluid ingested; indifference = 0.5] was computed and differences<br />

tested with analysis of variance (a =0.05). Hamsters strongly<br />

preferred sucrose over water [0.725]. Caf and Cyc aversions were<br />

unaffected by concentration or the addition of sucrose. Average<br />

preference ratios collapsed across concentration were 0.210 <strong>for</strong><br />

Caf and 0.225 <strong>for</strong> Caf + sucrose; 0.164 <strong>for</strong> Cyc and 0.151 <strong>for</strong><br />

Cyc + sucrose. Reducing bitter stimulus concentration did not<br />

increase amelioration by sucrose. This is consistent with Cyc<br />

aversions (1-hr tests) quickly developing, even to Cyc-sucrose<br />

mixtures without affecting sucrose intake (Hettinger et al. 2007),<br />

and Cyc serving as a CTA UCS when injected IP (Formaker<br />

et al. 2009). Acknowledgements: Supported by UConn SDM<br />

Alumni Research Fellowship and NIH grant DC004099.<br />

Regulation of Taste Responses by TNF<br />

Jinghua Chai, Pu Feng, Hong Wang<br />

Monell Chemical Senses Center Philadelphia, PA, USA<br />

Patients with inflammatory diseases often experience taste<br />

alterations. Yet, how inflammation affects taste function is<br />

not fully understood. Previously, we showed that tumor<br />

necrosis factor (TNF), a potent proinflammatory cytokine,<br />

is preferentially expressed in a subset of type II taste cells.<br />

The level of TNF in taste cells can be further induced by<br />

inflammatory stimuli such as lipopolysaccharide (LPS), a<br />

bacterial cell-wall component that elicits acute inflammation.<br />

Although TNF plays important roles in mediating inflammation<br />

and cell death in various tissues, its roles in taste buds remain<br />

to be determined. In this study, we carried out taste behavioral<br />

tests and gene expression analyses in wild-type and TNFdeficient<br />

mice. Lickometer tests were conducted to examine<br />

behavioral responses to salty, sour, bitter, sweet, and umami taste<br />

compounds be<strong>for</strong>e and after LPS-induced inflammation. Our<br />

results showed that TNF-deficient mice were less sensitive to<br />

the bitter compound quinine be<strong>for</strong>e any treatments. After LPS<br />

injection, wild-type mice displayed a range of altered responses<br />

to the taste compounds, especially to the sweet taste compound<br />

sucrose. In contrast, TNF-deficient mice did not show a<br />

significant alteration in response to sucrose after LPS treatment,<br />

suggesting that TNF plays an important role in regulating<br />

taste response to sucrose under LPS-induced inflammation.<br />

Furthermore, gene expression analyses by quantitative RT-PCR<br />

showed that the levels of several inflammation- and cell-deathrelated<br />

genes were increased by LPS in wild-type mice, but<br />

were not induced in TNF-deficient mice. Together, these results<br />

suggest that TNF may be an important mediator <strong>for</strong> taste<br />

dysfunction associated with inflammation. Acknowledgements:<br />

This study was supported by NIH/NIDCD grants DC010012<br />

and DC011735.<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

106


#P201 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

#P202 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

Evidence of neonatal memory of odor configuration<br />

Gérard Coureaud 1 , Thierry Thomas-Danguin 1 , Donald A. Wilson 2 ,<br />

Guillaume Ferreira 3<br />

1<br />

CSGA - Centre des <strong>Sciences</strong> du Goût et de l’Alimentation /<br />

Developmental Ethology and Cognitive Psychology Team & Flavour<br />

Perception Team Dijon, France, 2 Emotional Brain Institute / New York<br />

University School of Medicine New York, NY, USA, 3 Nutrition and<br />

Integrative Neurobiology (NutriNeuro) Group / INRA 1286, Université<br />

de Bordeaux Bordeaux, France<br />

The perception of some mixtures of odorants engages<br />

configural abilities, i.e. the perception of these mixtures as<br />

single odor objects. For instance, data in human adults<br />

demonstrated that a mixture of two odorants (AB), one smelling<br />

like strawberry and the other like caramel, generates the<br />

configural perception of the odor of pineapple (Le Berre et al.,<br />

2008; Barkat et al., 2012). Configural processing may be adaptive<br />

also <strong>for</strong> young organisms, to which rapid extraction of chemical<br />

in<strong>for</strong>mation from the maternal environment, highly complex,<br />

is a prerequisite to survival. Thus, results in newborn rabbits<br />

suggest the perception of a unique odor in the AB mixture<br />

(smelling like configural pineapple in humans) and different<br />

from the odors of the elements (Coureaud et al., 2008, 2009a).<br />

To clearly demonstrate that the configural AB perception does<br />

not directly depend on A and B perception, we investigated here<br />

whether rabbit neonates recognize the AB mixture even in the<br />

absence of A and B recognition. To that goal, rabbit pups were<br />

conditioned to AB on day 1. On day 2, recall of A and recall of<br />

B were followed by intraperitoneal injection of either saline or<br />

a pharmacological amnesic agent (see Coureaud et al., 2009b,<br />

2011). Testing <strong>for</strong> behavioral responsiveness to A, B and AB<br />

occurred on day 3. Control pups responded behaviorally to AB<br />

but also to A and B. As expected, the pups injected with the<br />

amnesic agent did not respond to A and to B. However, they<br />

responded to AB, indicating an AB perception independent of<br />

A and B representations. In summary, the present results confirm<br />

the perception by rabbit neonates of a configuration in the<br />

AB mixture, and demonstrate <strong>for</strong> the first time the neonatal<br />

ability to memorize odor mixtures as configurations independent<br />

of the memory of their elements. Acknowledgements: Supported<br />

by French ANR-2010-JCJC-1410-1 MEMOLAP to GC, TTD<br />

and GF.<br />

Sniffing Strategies in Wild-Type and Olfactory Marker<br />

Protein Knock-Out Mice<br />

Glen J. Golden 1 , Johannes Reisert 1 , Alan Gelperin 1,2<br />

1<br />

Monell Chemical Senses Center Philadelphia, PA, USA, 2 Princeton<br />

Neuroscience Institute, Department of Molecular Biology, Princeton<br />

University Princeton, NJ, USA<br />

Detection and identification of an odor requires nasal<br />

inhalation or sniffing behavior that delivers odorants to the<br />

olfactory receptor neurons (ORNs) deep in the nasal cavity. Our<br />

experiments combine behavioral assessment of odor detection<br />

and discrimination tasks with measurements of sniffing behavior<br />

to clarify the strategies a mouse uses when confronted with<br />

odor-based learning tasks and the mechanisms underlying odor<br />

perception. We study the behavior and sniffing patterns of mice<br />

with (WT) or without (KO) functional olfactory marker protein<br />

(OMP), a protein that is responsible <strong>for</strong> speeding up the time<br />

course of odor-induced responses in ORNs. OMP KO and WT<br />

mice were implanted with wireless pleural pressure sensors to<br />

record sniffing patterns. These mice were then trained and tested<br />

in go/no go odor discrimination tasks to distinguish solvent<br />

(mineral oil) odor from the odor of 1-propanol 10 -4 log dilution.<br />

Upon propanol or solvent exposure, WT mice increased their<br />

sniffing rate from ~4 Hz to 10 Hz and maintained a higher<br />

sniffing rate <strong>for</strong> rewarded (S+) trials. However, KO mice<br />

continued to increase their sniff rate following the onset of the<br />

odor (S+) and solvent-odor cues (S-) significantly in comparison<br />

to the WT mice (p


#P203 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

#P204 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

The Temporal Structure of Odor Mixture Perception in Rats<br />

Leslie M Kay 1,2 , Nisarg M Mehta 1 , Cinar Doruk 3<br />

1<br />

Institute <strong>for</strong> Mind and Biology, University of Chicago Chicago, IL,<br />

USA, 2 Department of Psychology, University of Chicago Chicago, IL,<br />

USA, 3 St. John’s College Annapolis, MD, USA<br />

Temporal structure of odor mixtures is often overlooked, but<br />

this is what gives many mixtures their complex percepts. This<br />

structure can help us understand the puzzling qualities of odor<br />

mixtures. There is not yet any theory that can predict whether a<br />

mixture of 2 monomolecular odorants will smell like both odors<br />

(elemental), like one more than the other (overshadowing), or<br />

like neither (configural or synthetic). One feature that is often<br />

overlooked in studies of mixture perception is the difference in<br />

perceptual arrival time <strong>for</strong> the two odors. These delays have been<br />

measured in humans, but implementing these sensitive perceptual<br />

assays in rats is much more difficult. We have developed a task<br />

that allows us to do this in rats, using a combined 2-alternative<br />

choice - go/no-go paradigm. The results show that behavioral<br />

response profiles to timing differences in binary mixtures are<br />

specific to the odor pair and that the responses to positive and<br />

negative delays are not symmetrical. For example, at zero delay<br />

(only the natural processes producing delays) in a 1:1 mixture of<br />

amyl acetate and anisole, anisole overshadows amyl acetate. As<br />

anisole is moved earlier in time overshadowing becomes stronger.<br />

In the negative direction, with amyl acetate preceding anisole, the<br />

mixture enters a configural regime at -50ms to -200ms and then<br />

takes on an elemental quality at -250ms. These results suggest<br />

that in the temporal domain, elemental and configural responses<br />

are close, with overshadowing responses occupying a separate<br />

part of the temporal space. The properties of odorants, such<br />

as sorptiveness and volatility, that may contribute to temporal<br />

effects are discussed. Acknowledgements: Institute <strong>for</strong> Mind and<br />

Biology Seed Grant (LK) Hodson Research Fellowship (CD)<br />

The Fine Temporal Structure of the Rat Licking Pattern:<br />

What Causes the Variability in the Interlick Intervals and<br />

How is it Affected by the Drinking Solution?<br />

Xiong B. Lin 1 , Dwight R. Pierce 2 , Kim E. Light 2 , Abdallah M. Hayar 1<br />

1<br />

University of Arkansas <strong>for</strong> Medical <strong>Sciences</strong>, Dept. of Neurobiology &<br />

Developmental <strong>Sciences</strong> Little Rock, AR, USA, 2 University of<br />

Arkansas <strong>for</strong> Medical <strong>Sciences</strong>, Dept. of Pharmaceutical <strong>Sciences</strong><br />

Little Rock, AR, USA<br />

Licking is a repetitive behavior controlled by a central pattern<br />

generator. Even though interlick intervals (ILI) within bursts<br />

of licks are considered fairly regular, the conditions that affect<br />

their variability are unknown. We analyzed the licking pattern<br />

in rats that were licking water, 10% sucrose solution, or 10%<br />

ethanol solution, in 90 min recording sessions after 4 h of water<br />

deprivation. The histograms of ILIs indicate that licking typically<br />

occurred at a preferred ILI of about ~135 ms with evidence of<br />

bi- or multi-modal distributions due to occasional licking failures.<br />

The longer the pause between bursts of licks (≥3 consecutive licks<br />

with ILIs 4 sec, the ILI<br />

was the shortest (~110 ms) at the beginning of the burst and then<br />

it increased rapidly in the first few licks and slowly in subsequent<br />

licks. The first ILI of a burst of licks was not significantly<br />

different when licking any of the 3 solutions, but subsequent<br />

licks exhibited a temporal pattern characteristic of each solution.<br />

Moreover, rats licked the ethanol solution in shorter bursts and<br />

the sucrose solution in longer bursts when compared to water.<br />

There<strong>for</strong>e, rats may rapidly identify the fluid and modify their<br />

licking behavior by adjusting the temporal pattern of licks and<br />

the number of licks/burst. The rapid deceleration in licking rate<br />

in bursts of licks that occurred after >4 sec pause was due to an<br />

increase from ~27 ms to ~56 ms in the tongue-spout contact<br />

duration while the intercontact interval was only slightly changed<br />

(80-90 ms). There<strong>for</strong>e, the contact duration seems to be the major<br />

factor that increases the variability in the ILIs, and could be<br />

another means <strong>for</strong> the rat to adjust the amount of fluid ingested.<br />

Acknowledgements: Grant P20 GM103425-09<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

108


#P205 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

#P206 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

How do rabbit newborns and human adults perceive the<br />

configuration in a 6-component blending odor mixture?<br />

Sébastien Romagny, Thierry Thomas-Danguin, Gérard Coureaud<br />

Centre des <strong>Sciences</strong> du Goût et de l’Alimentation (CSGA), UMR 6265<br />

CNRS, UMR 1324 INRA, Université de Bourgogne, Flavour Perception<br />

Team and Developmental Ethology and Cognitive Psychology Team<br />

Dijon, France<br />

Comparative studies give the opportunity to better understand<br />

common and dissimilar processes in terms of perception,<br />

cognition and behavior between organisms. Regarding the<br />

elemental and configural perception of odor mixtures, newborn<br />

rabbits and human adults present some similarities. In particular,<br />

they both process a 6-component mixture (RC) in a weak<br />

configural way, leading to the perception of a novel odor (e.g.,<br />

red cordial in humans) in the mixture: in rabbits, neonates do<br />

not respond to RC after learning of one component, while they<br />

generalize <strong>for</strong> another mixture of same complexity; in humans,<br />

the quality of the single odorants is judged as significantly<br />

different to the mixture. Here, we set out to examine whether<br />

the perception of the RC configuration strictly depends on the<br />

quantity and/or the quality of the components. To that goal,<br />

we carried out a generalization experiment in rabbit pups and<br />

a similarity rating task in human adults, using several submixtures<br />

of increasing complexity (i.e., 2, 3, 4 or 5 odorants). We<br />

conditioned the pups to sub-mixtures and tested their behavioral<br />

responsiveness to these stimuli compared to the full mixture<br />

(RC). In humans, participants rated the similarity between<br />

the sub-mixtures and RC. The results indicated that newborn<br />

rabbits became able to respond to the RC mixture when they had<br />

previously acquired at least 4 of its components, whatever their<br />

odor quality. In human adults, similarity between sub-mixtures<br />

and the RC mixture depended more on the odor quality of the<br />

odorants included in sub-mixtures rather than on the number of<br />

mixed odorants. There<strong>for</strong>e, even if these two models shared a<br />

configural perception of the same RC odor mixture, the factors<br />

underpinning its perception seem to be different, at least in part.<br />

The Contribution of the T1R1 Subunit to Taste Detection of<br />

Glutamate as Behaviorally Assessed in a Murine Model.<br />

Kimberly R Smith, Alan C Spector<br />

Florida State University Tallahassee, FL, USA<br />

Whether taste detection of L-glutamate is mediated solely<br />

by the T1R1+T1R3 heterodimer or whether an additional<br />

glutamate-sensing taste transduction mechanism(s) contributes<br />

is controversial. Here, we behaviorally assessed the necessity<br />

of the T1R1 subunit to taste detection of monosodium<br />

glutamate (MSG) in a two-response discrimination procedure<br />

using T1R1 knockout (KO) mice and their same-sex littermate<br />

wild-type (WT) controls. Water-restricted mice were trained<br />

to discriminate a tastant from water with a correct response<br />

resulting in the delivery of a water rein<strong>for</strong>cer and an incorrect<br />

response resulting in a time-out. Sensitivity to NaCl and MSG<br />

was similar between genotypes. However, upon the addition<br />

of the sodium channel blocker amiloride (100 µM) and inosine<br />

5’ monophosphate (IMP), at a concentration (2.5 µM) shown<br />

to potentiate the glutamate signal in a variety of assays,<br />

per<strong>for</strong>mance of the KO mice to this MSG mixture (M+A+I) was<br />

severely impaired. Whereas WT mice per<strong>for</strong>med at consistently<br />

high levels across concentrations, the ability of the KO mice to<br />

detect the M+A+I solution was above chance only at the higher<br />

MSG concentrations. The possibility that IMP was precluding<br />

concentration-dependent per<strong>for</strong>mance in the WT mice to the<br />

MSG in the presence of amiloride was confirmed when we<br />

found that WT mice could detect 2.5 µM IMP alone with<br />

relatively high accuracy, whereas KO mice could not respond<br />

significantly above chance. Collectively, these results strongly<br />

suggest that 1) the Na + ion dominates the taste detection of MSG<br />

in mice consistent with other recent data from our laboratory,<br />

and 2) glutamate may be activating a T1R1-independent highthreshold<br />

receptor in the presence of IMP, but normal detection<br />

of glutamate depends on the T1R1 subunit. Acknowledgements:<br />

NIH R01-DC004574 (ACS) & NSF GRF to KRS<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

109


#P207 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

#P208 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

A Role <strong>for</strong> Salivary Proteins in Taste Mediated Behavior<br />

Ann-Marie Torregrossa, Michelle B. Bales, Robert J. Contreras, James C.<br />

Smith , Lisa A. Eckel<br />

Dept. of Psychology, Florida State University Tallahassee, FL, USA<br />

While few studies have examined the influence of salivary<br />

proteins on taste-mediated behavior, it has been demonstrated<br />

that salivary proline-rich proteins (PRPs) bind to bitter-tasting<br />

tannic acid (TA) and may function to increase the acceptability<br />

of TA-containing diets. To test this hypothesis, we collected<br />

saliva samples and measured spontaneous feeding behavior<br />

in rats (n=8) fed a control diet (16 days) followed by a diet<br />

containing 3% TA (12 days). Total intake was reduced during<br />

the first 3 days of the TA diet (p


F344 rats on a two-odor discrimination task. Six odorants<br />

(1-propanol, 1-butanol, 1-pentanol, 1-hexanol, 1-heptanol and<br />

1-octanol) were arranged to produce 30 novel odorant pairs<br />

differing between one and five carbon atoms; testing sessions<br />

included presentation of only one randomly assigned pair daily<br />

(200 trials daily). Results showed that, although rats learned<br />

to discriminate between any two odorant pairs, discrimination<br />

accuracy changed systematically with carbon chain length<br />

difference. Error patterns were remarkably consistent across<br />

animals, such that marked increases in misses and false alarms<br />

were indicated <strong>for</strong> pairs differing by one or two carbon atoms.<br />

Across all odorant pairs, these effects were most pronounced<br />

during the first 20 trials. Notably, the greatest degree of<br />

perceptual confusion was displayed <strong>for</strong> two pairs differing by a<br />

single carbon atom, 1-propanol/1-butanol and 1-heptanol/1-<br />

hexanol. These data provide further support <strong>for</strong> carbon chain<br />

length as an important odorant stimulus dimension <strong>for</strong> study of<br />

olfactory receptor interaction (Johnson and Leon, 2000) as well<br />

as demonstrate how hierarchical chemotopic organization of<br />

the olfactory bulb may be reflected perceptually. Furthermore,<br />

development of an animal model using the carbon chain<br />

paradigm may be useful <strong>for</strong> assessing the mechanisms underlying<br />

olfactory dysfunction.<br />

#P210 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

The Gustatory Stop-Signal Task: A Method <strong>for</strong> Measuring<br />

Taste Quality Discrimination in Mice with Millisecond<br />

Temporal Resolution<br />

Dustin M Graham, David L Hill<br />

UVA/Psychology Charlottesville, VA, USA<br />

There is a lack of consensus regarding the roles of temporal and<br />

spatial coding of taste quality in the gustatory system due in<br />

part to various experimental and analytical differences among<br />

previous studies. Quantitative behavioral analysis can be used<br />

to test <strong>for</strong> the cognitive principle of speed-accuracy tradeoff<br />

(SAT), a hallmark of temporal processing of sensory stimuli.<br />

However, current methods used to study taste perception<br />

in rodents are temporally too slow <strong>for</strong> precise reaction-time<br />

measurements required to test <strong>for</strong> SAT in the gustatory system.<br />

We designed a novel behavioral paradigm, the Gustatory Stop-<br />

Signal Task, in head-restrained mice <strong>for</strong> measuring perceptual<br />

identification of taste stimuli with millisecond temporal<br />

resolution. Using this new paradigm, we will apply threshold<br />

psychophysics to determine if a stimulus-dependent SAT is<br />

present during discrimination of basic tastes. This will provide<br />

crucial behavioral evidence <strong>for</strong> the potential roles of temporal<br />

and spatial coding strategies underlying taste quality coding in<br />

the gustatory system. Additionally, the task can be combined<br />

with advanced physiological techniques, such as visually guided<br />

whole-cell patch-clamp recordings in sub-regions of gustatory<br />

cortex. The gustatory stop-signal task in head-restrained mice<br />

will provide a new foundation to combine precise quantitative<br />

behavioral measurements of taste perception alongside state-ofthe-art<br />

in vivo physiology. Acknowledgements: NIH Grants R01<br />

DC00407 and F32 DC012461 - 01A1<br />

#P211 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

Free Access to Highly Palatable Food during Adolescence<br />

Increases Anxiety- and Depression-like Behaviors in Males,<br />

but not in Females<br />

Jeong-Won Jahng, Jin-Young Kim, Joo-Young Lee, Jong-Ho Lee<br />

Seoul National University School of Dentistry Seoul, South Korea<br />

We have reported that a long-term access to highly palatable food<br />

(HPF) modulates the hypothalamic-pituitary-adrenal (HPA) axis<br />

response to restraint stress in adult male rats. Psycho-emotional<br />

disorders frequently involve dysfunctions in the HPA axis<br />

activity. In this study, male and female SD rats had free choices<br />

of chocolate cookies as HPF and chow with ad libitum access<br />

from PND 28, and then were subjected to behavioral tests at<br />

youth. Control group received chow only, and food conditions<br />

were continued throughout the whole experimental period. Body<br />

weight gain and daily caloric intake did not differ between HPF<br />

and control groups both in males and females. Total ambulatory<br />

activity was decreased with HPF access in females, but not in<br />

males. However, HPF increased anxiety related behaviors in<br />

males; i.e. increased rostral grooming and decreased the open<br />

arms stay during elevated plus maze test, but did not affect<br />

those indexes in females. Immobility duration during <strong>for</strong>ced<br />

swim test was significantly increased with HPF access in males,<br />

but the increase was not reached to a statistical significance in<br />

females. Stress-induced corticosterone increase was shortened<br />

with HPF access both in males and females. Results suggest that<br />

adolescence free access to highly palatable food may lead to a<br />

dysfunction in the HPA axis activity both in males and females;<br />

however, its psycho-emotional outcome be worse in males.<br />

Acknowledgements: Supported by MOEST(2009K001269 2010-<br />

0003642)<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

111


#P212 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

#P213 POSTER SESSION IV:<br />

CHEMICAL SIGNALING AND BEHAVIOR;<br />

ANIMAL BEHAVIOR/PSYCHOPHYSICS;<br />

CHEMOSENSATION AND METABOLISM;<br />

VOMERONSASAL AND CHEMICAL<br />

COMMUNICATION<br />

Licking Microstructure Reveals Rapid Attenuation of<br />

Neophobia<br />

Kevin J Monk, Benjamin D Rubin, Jennifer Keene, Donald B Katz<br />

Brandeis University Waltham, MA, USA<br />

Neophobia—the initial hesitation that many animals show to a<br />

novel food—is typically measured by comparing consumption<br />

in the first and second sessions of access to the taste; lower<br />

consumption in session 1 denotes neophobia, and higher 2ndsession<br />

consumption denotes attenuation of neophobia (AN).<br />

AN is thought to represent a bona fide example of learning—<br />

neural plasticity induced by an association between the taste and<br />

a safe outcome on session 1 changes the response to the tastant<br />

during session 2. Such long-term plasticity processes require time<br />

to complete, and thus AN should only stabilize 90 min or more<br />

following the first exposure to the tastant—a prediction that has<br />

been borne out in behavioral data. It remains possible, however,<br />

that a more rapidly developing AN might escape detection<br />

in time-averaged accounts of behavior such as consumption.<br />

With this in mind, we per<strong>for</strong>med a comparison of AN in two<br />

contexts, a two-bottle test and a brief access test (which allowed<br />

a real-time analysis of licking microstructure). At the level of<br />

overall consumption, data from the two tasks were in good<br />

accord—both revealed AN to 28mM saccharin but not to 2.8mM<br />

saccharin. Additionally, however, the brief access task revealed<br />

an initial hesitation to consume the higher concentration<br />

saccharin solution; this seemingly neophobia-related hesitation<br />

not only decreased between sessions 1 and 2, but also decreased<br />

linearly across the twenty minutes of session ; that is, AN began<br />

within minutes of the rats’ first exposure to the taste. These<br />

data validate the brief-access task as an paradigm with which to<br />

measure AN, and also reveal aspects of AN—perhaps related<br />

to short-term plasticity—that appear within minutes of the first<br />

taste. Acknowledgements: National Institutes of Health, World<br />

of Work Fellowship<br />

Channelrhodopsin Mice use Temporal In<strong>for</strong>mation Encoded<br />

in the Olfactory Bulb <strong>for</strong> Odor Sensation.<br />

Michelle R. Rebello 1,2 , Thomas S. McTavish 2 , David C. Willhite 1,2 ,<br />

Gordon M. Shepherd 2 , Justus V. Verhagen 1,2<br />

1<br />

John B. Pierce Laboratory New Haven, CT, USA, 2 Yale School of<br />

Medicine, Dept. of Neurobiology New Haven, CT, USA<br />

Odor in<strong>for</strong>mation is represented by spatio-temporal maps in the<br />

olfactory bulb (OB). Spatial maps reflect the converging axons<br />

of olfactory receptor neurons activated by odors, onto their<br />

respective glomeruli in the OB. The origins of temporal patterns<br />

of glomerular activation are less well understood, but odorant<br />

receptor affinity as well as odorant sorption kinetics across<br />

the olfactory epithelium could underlie temporal parameters<br />

such as onset latency and rise time. Consistent differences in<br />

response dynamics across glomeruli have been found <strong>for</strong> odorevoked<br />

responses in the OB. Further, we have shown, using<br />

optical imaging that retronasal and orthonasal bulbar reponses<br />

differ in response amplitude as well as temporal dynamics. It<br />

is there<strong>for</strong>e evident that rich temporal in<strong>for</strong>mation is available<br />

in the bulbar response. However it is not known whether these<br />

temporally dynamic responses are behaviorally relevant. Using<br />

transgenic mice expressing ChR2 under the Thy-1 promoter in<br />

the mitral cells and a digital micromirror device to project snifftriggered<br />

light patterns onto the dorsal OB we are able to exert<br />

tight spatio-temporal control over OB activity patterns. We find<br />

that mice trained on a go/no-go task are able to discriminate<br />

patterns that are spatially identical but differ temporally. By<br />

varying the relative delay among the same regions activated by<br />

light patterns we are able to determine the threshold of temporal<br />

discrimination. We find that Thy-1 ChR2 but not wild-type<br />

mice can make temporal discriminations of less than 30ms.<br />

Our optogenetic study confirms that awake, behaving mice<br />

can use temporal in<strong>for</strong>mation encoded in the bulbar response.<br />

This suggests that temporal coding can contribute to retronasal<br />

and orthonasal odor sensation. Acknowledgements: This work<br />

is supported by NIH/NIDCD Grants R01DC009994 and<br />

R01DC011286 and NIH Institutional Training Grant T15-<br />

LM007056 from the National Library of Medicine.<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

112


#P214 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

Potentiation of Primary Afferent Innervation in the<br />

Rostral Nucleus of the Solitary Tract<br />

Robert M. Bradley, James A. Corson<br />

University of Michigan/ Biologic and Materials <strong>Sciences</strong> Ann Arbor,<br />

MI, USA<br />

The development of mature primary afferent circuitry results<br />

from a Hebbian-like synaptic strengthening in a number of<br />

sensory systems. This synaptic strengthening is accompanied<br />

by a pruning of non-strengthened inputs resulting in a mature<br />

terminal field anatomy. Gustatory afferents innervating the oral<br />

cavity project to the rostral nucleus of the solitary tract (rNTS),<br />

the terminal fields of which prune between postnatal days 15<br />

and 60 into an adult-like organization. However, it is not known<br />

whether any activity dependent changes in synaptic strength<br />

can be induced during this period that may correspond to the<br />

anatomical remodeling. To investigate the activity-modulated<br />

plasticity of primary afferent inputs to the rNTS, acute<br />

horizontal rNTS slices were prepared from rats of postnatal ages<br />

covering the period of anatomical plasticity. The solitary tract<br />

was stimulated with a concentric bipolar electrode and excitatory<br />

postsynaptic currents (ePSC) were recorded. The ability of<br />

a synapse to be potentiated was investigated by stimulating<br />

the solitary tract at 50 Hz paired with a 5 ms depolarization.<br />

Following this tetanic stimulation increases in ePSC amplitude<br />

and rise slope were observed in each age examined, though only<br />

in a subset of neurons at each age. Interestingly, the probability<br />

of inducing potentiation appears to decrease between postnatal<br />

days 20-25 and subsequently increases from postnatal day 30<br />

onward. Primary afferent potentiation lasted over variable<br />

time scales ranging from 1 to 30 minutes, indicative of either<br />

short-term or long-term potentiation. These results suggest that<br />

subpopulations of primary afferent synapses can be potentiated<br />

throughout development, and the presynaptic nerve and/<br />

or postsynaptic target specificity <strong>for</strong> this potentiation will the<br />

focus of future studies. Acknowledgements: T32DC000011,<br />

RO1DC000288<br />

#P215 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

Ephrin-B/EphB signaling influences the innervation of<br />

fungi<strong>for</strong>m papillae<br />

David Collins 1 , Natalia Hoshino 1 , Elizabeth M Runge 1 , Son Ton 1 ,<br />

Omar Diaz 1 , Jessica Decker 1 , Mark Henkemeyer 2 , M William Rochlin 1<br />

1<br />

Loyola U Chicago/Biology Chicago, IL, USA, 2 UT Southwestern/<br />

Developmental Biology Medical Center/ Dallas, TX, USA<br />

Geniculate ganglion axons innervate fungi<strong>for</strong>m taste buds<br />

whereas the trigeminal ganglion neurites innervate the adjacent<br />

non-taste epithelium. Owing to the proximity of their terminal<br />

fields, non-diffusible repellents are likely to have a role in<br />

preventing incorrect targeting. Eph receptors and ephrins are<br />

cell-attached receptor/ligands capable of triggering contactdependent<br />

repulsion or stabilization. An antibody that detects<br />

EphB1, B2, and B3 stains both geniculate and trigeminal axons<br />

throughout embryonic development. Anti-ephrin-B1 and antiephrin-B2<br />

labeled the dorsal lingual epithelium uni<strong>for</strong>mly at<br />

E17, when axons first penetrate the papilla epithelium in rat.<br />

In ephrin-B2 lacZ mice, label was also observed throughout the<br />

dorsal epithelium but only from E16.5, well after initial invasion<br />

of axons into the epithelium (E14.5). However, in ephrin-B1<br />

lacZ mice, only fungi<strong>for</strong>m papilla epithelium was labeled, and<br />

this labeling was observed at E14.5. Mice lacking EphB1 and<br />

EphB2 exhibited normal levels of gustatory innervation at<br />

E13.5, but less innervation than controls by E17.5, suggesting<br />

that ephrin-B/EphB <strong>for</strong>ward signaling may have a stabilizing<br />

influence on gustatory afferents in vivo. In vitro, substratum<br />

stripes prepared from high concentrations of ephrin-B2-Fc (40<br />

ug/ml) repel neurites from both the geniculate and trigeminal<br />

ganglia, and this did not depend on which neurotrophin<br />

was used to promote growth. Stripes prepared from lower<br />

concentrations of ephrin-B2-Fc (4 ug/ml) were not repellent;<br />

indeed, coverglasses coated uni<strong>for</strong>mly with 4 ug/ml ephrin-B2<br />

mildly promoted trigeminal neurite outgrowth length, depending<br />

on the stage and neurotrophin. We are currently analyzing<br />

the effects of ephrin-B1 on geniculate and trigeminal neurites.<br />

Acknowledgements: 1R15DC010910-01<br />

#P216 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

Glial Contributions to the Formation of the Solitary Tract and<br />

the Rostral Nucleus of the Solitary Tract<br />

Sara L Corson, Robert M Bradley, Charlotte M Mistretta<br />

University of Michigan School of Dentistry Department of Biologic and<br />

Materials <strong>Sciences</strong> Ann Arbor, MI, USA<br />

The solitary tract (ST) consists of afferent fibers that originate in<br />

the oral cavity and project to the rostral nucleus of the solitary<br />

tract (rNST), the site of the first synaptic relay in transmitting<br />

taste-related in<strong>for</strong>mation to higher brain areas. We are interested<br />

in the regulatory elements that direct ST <strong>for</strong>mation and rNST<br />

development. Glia represent approximately half of the cells<br />

in the CNS and play roles in neuron guidance and synapse<br />

development and function. However, the time course of glial<br />

development and the role of glia in the gustatory brainstem are<br />

unknown. We surveyed the expression of glial and neuronal<br />

markers in the pre- and post-natal developing rat ST and rNST<br />

to characterize their contribution to the development of the<br />

gustatory brainstem. We examined the expression of neuronal<br />

markers, including calbindin and NeuN, and glial markers,<br />

including glial fibrillary acidic protein (GFAP), myelin basic<br />

protein (MBP) and brain lipid binding protein (BLBP), in<br />

conjunction with P2X2, a marker of gustatory nerve terminal<br />

fields, in the developing ST and rNST. We found persistent but<br />

dynamic expression of calbindin, GFAP and BLBP throughout<br />

pre- and post-natal development. In particular, GFAP expression<br />

shifts from more fibrillary to more astrocyte-like labeling in the<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

113


early postnatal period. This shift is concurrent with terminal<br />

field plasticity, i.e., developmental remodeling and fine-tuning<br />

of circuits. Furthermore, the GFAP-positive astrocyte-like<br />

cells are differentially distributed throughout the rNST. This<br />

work highlights potential neuron-glia interactions that are<br />

important <strong>for</strong> the development of the gustatory brainstem.<br />

The results provide basic in<strong>for</strong>mation <strong>for</strong> building mechanistic<br />

studies regarding glial function in taste circuit <strong>for</strong>mation.<br />

Acknowledgements: NIH NIDCD Grant DC009982<br />

#P218 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

The p75 receptor regulates gustatory innervation patterns<br />

during development<br />

Da Fei, Robin Krimm<br />

University of Louisville/Anatomical <strong>Sciences</strong> and Neurobiology<br />

Louisville, KY, USA<br />

#P217 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

Functions of the GDNF family of neurotrophic factors in the<br />

development of the peripheral gustatory system<br />

Christopher R. Donnelly, Brian A. Pierchala<br />

University of Michigan School of Dentistry/Biologic and Materials<br />

<strong>Sciences</strong> Ann Arbor, MI, USA<br />

During development of the peripheral nervous system (PNS),<br />

target-derived neurotrophic factors, such as the neurotrophins<br />

and the glial cell-line derived neurotrophic factor (GDNF) family<br />

ligands (GFLs), are critical <strong>for</strong> the establishment of proper<br />

connections between neurons and their targets. The four GFLs,<br />

GDNF, neurturin (NRTN), artemin (ARTN) and persephin<br />

(PSPN), are potent growth, guidance and survival factors<br />

<strong>for</strong> both PNS and CNS neurons. The GFLs bind with high<br />

affinity to GPI-anchored coreceptors called the GFRalphas, of<br />

which there are four (GFRalpha1-4). In the PNS, these GFL-<br />

GFRalpha complexes then associate with and activate their<br />

common receptor tyrosine kinase, Ret. Several of the GFLs and<br />

their receptors are expressed by components of the peripheral<br />

gustatory system. GDNF and NRTN are expressed throughout<br />

the lingual epithelium, and GDNF, GFRalpha1, GFRalpha2<br />

and Ret are all expressed in taste buds of circumvallate papillae.<br />

In addition, neurons of the petrosal and geniculate ganglia<br />

express Ret and cognate GFRalphas. It is not known, however,<br />

whether GFLs function in the development and maintenance<br />

of peripheral gustatory circuits. Using conditional transgenic<br />

deletion of Ret, we are in the process of determining whether the<br />

GFL/Ret signaling pathway is necessary <strong>for</strong> the development<br />

of fungi<strong>for</strong>m papillae, taste buds within circumvallate and<br />

fungi<strong>for</strong>m papillae, and their respective innervation by petrosal<br />

and geniculate neurons. These studies will establish whether<br />

other neurotrophic factors besides BDNF and NT-4 have<br />

developmental functions in the embryonic morphogenic events<br />

in the lingual epithelium necessary <strong>for</strong> papillae and taste bud<br />

development, as well as growth and survival functions <strong>for</strong><br />

the sensory neurons that innervate these lingual structures.<br />

Acknowledgements: BAP: NINDS R01 NS058510 CRD:<br />

NIDCR TEAM Tissue Engineering and Regeneration Training<br />

Grant T32 DE007057<br />

As a pan receptor of neurotrophins, p75 can function as either<br />

a pro-survival or pro-death factor during development. P75 is<br />

expressed in both taste buds and taste (geniculate) neurons;<br />

however, the role of p75 receptor during taste development is<br />

unclear. Here we examined the role of p75 in neuron survival,<br />

taste bud <strong>for</strong>mation and peripheral innervation patterns during<br />

development. We found that p75 -/- mice began to lose about<br />

22% (p=0.05) of geniculate neurons compared to the wild type<br />

mice at E14.5, and the loss continued to around 36% (p


fide adult stem cell marker. Because canonical Wnt signaling<br />

is activated at sites of mature taste buds, we asked if Lgr5<br />

expression identifies analogous stem/progenitor cells in the<br />

taste system. Knock-in mice with the Lgr5-EGFP-ires-CreERT2<br />

transgene replacing one Lgr5 allele were bred to the tamoxifen<br />

(TM)-inducible mTmG reporter strain. The progeny express<br />

soluble GFP (sGFP) under endogenous Lgr5 regulation and,<br />

following TM treatment, membrane-bound tdTomato (mT)<br />

is replaced by GFP (mG). sGFP is present in all papillae of<br />

7 day-old non-induced mice, but by 12 weeks of age is seen<br />

only in circumvallate (CV) and foliate papillae. Expression in<br />

the CV is highest where salivary ducts intersect papilla walls,<br />

decreases dorsally in the papilla, and is absent from taste buds.<br />

In salivary ducts, sGFP is limited to the outer layer of the bilayered<br />

epithelium. One day after a single TM injection, induced<br />

mG marks small round cells at the papilla base and, rarely, in<br />

the lateral wall of the CV papilla. By 2-3 days post-injection,<br />

labeled cells lie both within taste buds and in the surrounding<br />

stratified epithelium. Although mG-positive cell numbers decline<br />

by 7 days, induced labeling is still apparent in taste buds after<br />

2 months. Importantly, we found no evidence <strong>for</strong> TM-induced<br />

labeling of cells outside of taste papilla boundaries. These<br />

data indicate that Lgr5 expression marks long-term taste cell<br />

progenitors, that Lgr5 may be a regulator of taste epithelium<br />

homeostasis, and that a duct-basal trench-papilla axis of<br />

epithelial renewal may exist. Acknowledgements: Supported by<br />

NSF REU #1062645<br />

#P220 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

Temporal and spatial differences in BDNF and NT4 expression<br />

determine their unique roles in gustatory development<br />

Tao Huang 1 , Robin Krimm 1<br />

1<br />

University of Louisville/Department of Anatomical <strong>Sciences</strong> and<br />

Neurobiology Louisville, KY, USA, 2 University of Louisville/<br />

Department of Anatomical <strong>Sciences</strong> and Neurobiology Louisville,<br />

KY, USA<br />

A limited number of growth factors are capable of regulating<br />

numerous developmental processes, but how they accomplish<br />

this is unclear. In the gustatory system, brain-derived<br />

neurotrophic factor (BDNF) and neurotrophin-4 (NT4) have<br />

different developmental roles but exert their effects through<br />

the same receptors (TrkB and p75). Using genome wide<br />

expression analysis, we determined that BDNF and NT4<br />

regulate the expression of different sets of genes downstream<br />

of receptor signaling in gustatory ganglion. These differences<br />

in gene expression likely determine their different roles during<br />

development. BDNF and NT4 could function differently<br />

because of temporal or spatial differences of expression or the<br />

activation of different signaling pathways. Using mice in which<br />

the coding region <strong>for</strong> BDNF is replaced with NT4 (Bdnf Nt4/<br />

Nt4<br />

), we show that NT4 can mediate most of the unique roles<br />

of BDNF. Specifically, caspase-3-mediated cell death, which<br />

is increased in Bdnf -/- mice (p


#P222 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

#P223 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

Multiple Shh Signaling Centers in Embryo and Adult<br />

Participate in Fungi<strong>for</strong>m Papilla and Taste Bud Formation<br />

and Maintenance<br />

Hongxiang Liu 1 , Alex Ermilov 2 , Marina Grachtchouk 2 , Libo Li 1 ,<br />

Deborah L Gumucio 3 , Andrezej A Dlugosz 2,3 , Charlotte M Mistretta 1<br />

1<br />

Department of Biologic and Materials <strong>Sciences</strong>, School of Dentistry,<br />

University of Michigan Ann Arbor, MI, USA, 2 Department of<br />

Dermatology, Medical School, University of Michigan Ann Arbor, MI,<br />

USA, 3 Department of Cell and Developmental Biology, Medical School,<br />

University of Michigan Ann Arbor, MI, USA<br />

Fungi<strong>for</strong>m papillae must contain long-lived sustaining cells<br />

and short-lived maintaining cells that support development,<br />

differentiation and maintenance of the lateral and apical papilla<br />

epithelium and the specialized taste buds. Shh is a known<br />

regulator of papilla development but details about locations<br />

of ligand, target responding cells and transcriptional activators<br />

<strong>for</strong> Shh signaling are not known. We used immunostaining, in<br />

situ hybridization and reporters <strong>for</strong> Shh, Ptch1, Gli1 and Gli2-<br />

expressing cells to identify proliferating and differentiating cells<br />

in embryonic, postnatal and adult tongue, in papilla placodes,<br />

fungi<strong>for</strong>m papillae and/or taste bud cells that participate in<br />

Shh signaling. Whereas there is a progressive restriction in<br />

location of the Shh ligand, a receptive surround of Ptch1 and<br />

Gli1 expression in responding cells is maintained in particular<br />

epithelial and mesenchymal signaling centers throughout<br />

papilla development and taste bud differentiation. From lineage<br />

tracing, we know that Gli1-expressing cells and their progeny are<br />

located in fungi<strong>for</strong>m papilla basal cells, in perigemmal cells and<br />

mesenchymal cells of the papilla core, and are progenitors of<br />

taste cells. Further, using a doxycycline-regulated bitransgenic<br />

GLI2* mouse, in a functional test of activated Shh signaling<br />

in postnatal tongue epithelium, there is loss of fili<strong>for</strong>m papilla<br />

spines and loss of fungi<strong>for</strong>m papillae and taste buds. Loss of<br />

papilla organs is accompanied by proliferation in suprabasal<br />

layers of the lingual epithelium. The synthesized data position<br />

Shh signaling in multiple centers that are essential to placode<br />

and papilla development, and to postnatal papilla and taste bud<br />

differentiation and maintenance. Shh roles are most likely via<br />

paracrine mechanisms, and engage epithelial/mesenchymal<br />

interactions. Acknowledgements: NIH Grants NIDCD<br />

DC000456 (CMM), NIDDK DK065850 (DLG), NCI CA087837<br />

(AAD).<br />

BDNF is Required <strong>for</strong> the Development of Adult Taste Bud<br />

Number and Normal Behavioral Responses to Sour Stimuli<br />

Abigail B. Menefee 1 , Robin F. Krimm 2<br />

1<br />

dupont Manual High School Louisville, KY, USA, 2 Dept. of<br />

Anatomical <strong>Sciences</strong> and Neruobiology, Univeristy of Louisville Medical<br />

School Louisville, KY, USA<br />

Brain derived neurotrophic factor (BDNF) regulates gustatory<br />

system development. Because BDNF removal is neonatal lethal,<br />

the long-term effects of BDNF removal on the structure and<br />

function of the adult gustatory system are unclear. To address<br />

this issue we examined the adult taste system in conditional<br />

Bdnf knockouts in which Bdnf expression is reduced to one-tenth<br />

normal levels in the entire animal (Bdnf lox/lox ) and is completely<br />

removed from the lingual epithelium (K14-Cre;Bdnf lox/lox ).<br />

K14-Cre;Bdnf lox/lox mice had very few fungi<strong>for</strong>m taste buds<br />

remaining (11 ± 2) compared to wild type (52 ± 5, p ≤ 0.002)<br />

or Bdnf lox/lox mice (42 ± 10; p ≤ 0.02). The K14-Cre;Bdnf lox/lox<br />

circumvallate papillae contained 25% fewer taste buds than the<br />

control genotypes (p ≤ 0.025). There was no difference in taste<br />

bud number between wild type and Bdnf lox/lox , even though Bdnf<br />

lox/lox<br />

mice have substantially reduced Bdnf expression. There<strong>for</strong>e,<br />

as long as some BDNF remains, normal taste bud numbers<br />

are maintained. Short-term lick rate tests of K14-Cre;Bdnf lox/<br />

lox<br />

, Bdnf lox/lox , and wild type mice were used to examine taste<br />

function. Surprisingly, in spite of the large reduction in taste bud<br />

number, there was no statistical difference among the genotypes<br />

in lick rates to sucrose, quinine, and NaCl. This indicates that<br />

normal behavioral taste responses can be maintained in mice<br />

with few fungi<strong>for</strong>m taste buds. However, K14-Cre;Bdnf lox/lox mice<br />

have higher lick rates to citric acid at pH=3.2 (p ≤ 0.024) and<br />

pH=2.8 (p ≤ 0.01) compared to wild type mice. This indicates<br />

that removal of BDNF may cause a specific deficit in sour taste,<br />

which cannot be explained simply by the loss of taste buds.<br />

Acknowledgements: DC007176<br />

#P224 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

Reorganization of Primary Afferent Terminal Fields in the<br />

Mouse Brainstem Produced by Early Prenatal Dietary Sodium<br />

Restriction<br />

Chengsan Sun, David L Hill<br />

University of Virginia/Psychology Charlottesville, VA, USA<br />

Age-related decreases in terminal field volumes of the rat GSP,<br />

CT, and IX nerves and their overlapping fields in the nucleus<br />

of the solitary tract (NST) occur during normal development.<br />

The processes involved in “pruning” the three terminal fields<br />

can be altered significantly when rats are fed a sodium-restricted<br />

diet from E3-E12. All terminal fields are relatively large during<br />

early postnatal ages and thereafter fail to “prune”. Surprisingly,<br />

many of the terminal fields in restricted rats expand after 35 days<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

116


postnatal. Thus, a very early period of dietary sodium restriction<br />

leads to a late-onset expansion of terminal fields in the rat<br />

NST. To begin identifying the cellular/molecular mechanisms<br />

responsible <strong>for</strong> this brainstem plasticity, we explored the terminal<br />

field organization in adult mice that either received a sodiumreplete<br />

diet throughout development (controls) or mice fed the<br />

sodium-deficient diet from E3-E12. Moreover, we counted the<br />

number of ganglion cells, representing the three nerves, recorded<br />

whole-nerve neurophysiological taste responses from the CT and<br />

IX, and conducted 48hr. 2-bottle preference tests to concentration<br />

series of NaCl, sucrose, quinine, and citric acid. Terminal field<br />

volumes <strong>for</strong> each nerve were significantly greater (2X – 4X)<br />

in early sodium-restricted mice, and the overlapping zone that<br />

received all three nerves was over 15X greater in restricted mice.<br />

The differences in terminal fields were accompanied by increased<br />

preferences to NaCl and decreased aversions to citric acid, no<br />

differences in CT and IX whole-nerve taste responses, and no<br />

differences in number of GSP, CT, and IX ganglion cells. We<br />

conclude that the early dietary manipulation had a profound<br />

effect on early NST development and that the terminal field<br />

alterations impacted taste-related behaviors. Acknowledgements:<br />

R01 DC00407<br />

#P225 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

Renewal Kinetics of Taste Bud Cells in Adult Mice<br />

Linda A. Barlow 1,2 , Brendan W. Ross 1,2 , Lauren A. Gross 1,2<br />

1<br />

Dept of Cell & Developmental Biology, University of Colorado School<br />

of Medicine Aurora, CO, USA, 2 Rocky Mountain Taste & Smell Center,<br />

University of Colorado School of Medicine Aurora, CO, USA<br />

Taste bud cells are continuously renewed from a population<br />

of progenitor cells, which express cytokeratin (K)5. These<br />

progenitors reside outside taste buds and give rise to daughter<br />

cells, which exit the cell cycle, enter buds and differentiate into<br />

one of 3 taste cell types (I, II, III). Differentiated taste cells live,<br />

on average, <strong>for</strong> 10 days, although variance around this mean is<br />

large, suggesting different cell types have different longevities<br />

(Beidler & Smallman ‘65 J Cell Biol 27:263). To explore the idea<br />

of cell type-specific lifespans, we employed inducible genetic<br />

birthdating, comprising a K5rtTA driver and a tetracyclineinducible<br />

reporter allele, which encodes a nucleosomal protein,<br />

Histone2B fused with GFP (tetO-H2BGFP, Tumbar et al.,<br />

2004 Science 303:359). When K5rtTA;tetO-H2BGFP mice eat<br />

doxycycline (dox) chow, K5+ cells produce H2BGFP, which is<br />

incorporated into nucleosomes during S phase. Once mice are<br />

taken off dox, H2BGFP is no longer transcribed, and the cohort<br />

of GFP labeled cells comprises the “pulse”. Here, bigenic mice<br />

were fed dox chow <strong>for</strong> 12 hours, and tongues harvested between<br />

3 and 28 days. In the posterior circumvallate papilla, an average<br />

of 2 cells/ bud was GFP+ after a 3 day chase. This increased to 4<br />

GFP+ cells/bud by 14 days, and persisted through 21 days postdox.<br />

At 28 days, however, only 2 cells/bud were GFP+. At 14<br />

days, 2 GFP+ cells/bud were PLCß2+ type II cells, whereas on<br />

average, less than 1 GFP+ cell/ bud were Snap25+ type III cells.<br />

The number of type II cells/bud began to decline at 21 days,<br />

while type III cells/bud were fewer by 28 days. We are currently<br />

determining if type II cell lifespan is less than that of type III<br />

cells, and/or if type III cells are generated less frequently from<br />

later divisions of GFP+ K5+ progenitors. Acknowledgements:<br />

R01 DC012383 to LAB P30 DC004657 to D. Restrepo<br />

#P226 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

Bitter taste similarities among heterozygous MZ twins<br />

compared with homozygous MZ twins.<br />

Suzie Alarcon 1 , Ashley Sharples 1 , Paul A. S. Breslin 1,2<br />

1<br />

Rutgers, The State University of New Jersey New Brunswick, NJ,<br />

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

Heterozygous alleles of select genes are known to be expressed<br />

differentially in different individuals. And <strong>for</strong> some genes,<br />

expression in heterozygous monozygotic (MZ) twins is<br />

under strong genetic control (twins resemble each other in<br />

allele expression, but differ from twin pair to twin pair).<br />

The mechanism of action of differential allelic expression is<br />

unclear. In order to investigate the potential genetic influence<br />

of differential expression of taste genes, we examined the bitter<br />

taste phenotypic response of MZ (identical) twin pairs to the<br />

bitter tastant 6n -propyl-2-thiouracil (PROP). Volunteers at the<br />

Twinsburg Twins Days Festival in Cleveland OH tasted PROP<br />

and recorded their perceived bitterness intensity on a generalized<br />

labeled magnitude scale (LMS). DNA samples were collected by<br />

cheek swab from each subject and were genotyped at TAS2R38<br />

amino acid codon positions 49, 262, and 296. MZ twins<br />

were grouped by their TAS2R38 diplotype. We compared the<br />

similarity of Twin A vs Twin B <strong>for</strong> heterozygous MZ twins to the<br />

similarity of Twin A vs Twin B <strong>for</strong> homozygous MZ twins. The<br />

heterozygous MZ group displayed greater perceptual variation<br />

between twins than the homozygous MZ group, despite the fact<br />

that the homozygous group was comprised of both AVI/AVI<br />

and PAV/PAV homozygous subgroups. Thus, to the degree that<br />

phenotype reflects expression, it appears that allelic expression<br />

of bitter taste receptors is under significant non-genetic control<br />

within heterozygous MZ twins. Acknowledgements: Funded in<br />

part by NIH DC 02995 to PASB.<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

117


#P227 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

#P228 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

TAS2R polymorphisms and the bitterness of RebaudiosideA<br />

and RebaudiosideD<br />

Alissa L. Allen 1 , John E. McGeary 2 , John E. Hayes 1<br />

1<br />

Department of Food Science, Penn State University Park, PA, USA,<br />

2<br />

Providence VA Medical Center Providence, RI, USA<br />

With growing demand <strong>for</strong> natural non-nutritive sweeteners,<br />

extracts from the Stevia rebaudiana Bertoni plant have become a<br />

popular replacement <strong>for</strong> sugar. These exacts contain a mixture<br />

of various taste active glycosides, with steviol being the most<br />

abundant. The second most abundant glycoside is Rebaudioside<br />

A (RebA). RebA (>98% pure) is a GRAS (generally recognized<br />

as safe) ingredient in the United States, while stevia (the<br />

mixed extract) is classified as a dietary supplement. Another<br />

glycoside of interest is rebaudioside D (RebD), due to its high<br />

dose response <strong>for</strong> sweetness and minimal bitterness relative to<br />

steviol and RebA. Like saccharin and acesulfameK (AceK), the<br />

bitterness from stevia glycosides varies across people. Previously,<br />

variable saccharin/AceK bitterness has been associated with<br />

single nucleotide polymorphisms (SNPs) in bitter receptor<br />

genes (TAS2Rs). As part of an ongoing study, we explored<br />

whether TAS2R SNPs may explain variable RebA and RebD<br />

bitterness. After a brief orientation with bitter, sweet and metallic<br />

training references, participants rated RebA, RebD, aspartame,<br />

sucrose, and gentiobiose (a b 1-6 linked disaccharide) <strong>for</strong> these<br />

sensations on a general Labeled Magnitude Scale. Salivary DNA<br />

was obtained and genotyped via Sequenom MassARRAY. As<br />

expected, mean RebD bitterness was much lower than RebA. In<br />

preliminary analyses, RebA bitterness associated with a coding<br />

SNP in TAS2R9 and a synonymous SNP in TAS2R50. For the<br />

TAS2R9 Ala187Val SNP, Ala187 allele carriers reported less<br />

bitterness than Val187 homozygotes. For RebD, the TAS2R50<br />

SNP predicted bitterness, although this is likely a tag SNP <strong>for</strong><br />

another polymorphism given the synonymous substitution.<br />

Finally, TAS2R31 SNPs previously shown to predict AceK<br />

and saccharin bitterness did not predict bitterness from RebA<br />

or RebD. Acknowledgements: Supported by funds from the<br />

Pennsylvania State University and NIH grant DC0010904.<br />

Sucrose analgesia and the cold pressor test in young men:<br />

Methodological considerations<br />

Rachel G Antenucci 1,2 , John Prescott 3 , Theresa L White 4 ,<br />

John E Hayes 1,2<br />

1<br />

Department of Food Science, The Pennsylvania State University<br />

University Park, PA, USA, 2 Sensory Evaluation Center, The<br />

Pennsylvania State University University Park, PA, USA, 3 TasteMatters<br />

Research & Consulting Sydney, Australia, 4 Department of Psychology,<br />

LeMoyne College Syracuse, NY, USA<br />

Sucrose is mildly analgesic in infant rats, human neonates, and<br />

prepubertal children. This effect generalizes to non-nutritive<br />

sweeteners, indicating the effect is due to sweet taste. However,<br />

consistent sweet analgesia in adults remains elusive: some studies<br />

find an effect while others do not. Pain can be safely induced<br />

in the laboratory using the Cold Pressor Test (CPT), providing<br />

a convenient means to study this phenomenon. It is unresolved<br />

whether the inability to consistently observe analgesic effects in<br />

adults is due to methodological issues associated with the CPT<br />

or an age dependent effect. White and Prescott (AChemS, 2012)<br />

reported strong order effects; tolerance was greater in the second<br />

CPT within a session. They suggested rewarming the hand<br />

between tests might reduce this bias. Here, we describe 2 studies<br />

that attempt to refine an adult sweet analgesia CPT paradigm.<br />

In study 1 (36 men), pain was induced by placing the dominant<br />

hand in circulating water at 8 o C. Within a session, tastants were<br />

water and 0.7M sucrose; participants were tested twice (fed /<br />

fasted) in a double crossover design. The hand was rewarmed in<br />

35 o C water between trials, and hand temperature was confirmed<br />

with a laser thermometer. No tastant effect, regardless of hunger<br />

state, was observed. However, a weak trend <strong>for</strong> an order effect<br />

within a session was evident, despite rewarming. In study 2 (27<br />

men), subjects received sucrose or water on separate days after an<br />

overnight fast. The cold bath was reduced to 4 o C, and outcomes<br />

included pain threshold (onset in s) and tolerance (hand<br />

withdrawl in s). Pain thresholds and tolerance were greater on<br />

day 2, but this occurred irrespective of tastant. Collectively, these<br />

data suggest increased tolerance in the CPT with repeat exposure<br />

is not an effect of initial hand temperature. Acknowledgements:<br />

The Pennsylvania State University and USDA Hatch Act funds<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

118


#P229 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

Are individuals with elevated food liking scores (‘foodies’)<br />

hypergeusic?<br />

Nadia K Byrnes, Alissa L Allen, Emma L Feeney, Rachel J Primrose,<br />

John E Hayes<br />

Department of Food Science University Park, PA, USA<br />

The public believes that foodies are supertasters (and vice versa),<br />

consistent with reports that chefs and wine experts report greater<br />

bitterness from propylthiouracil (PROP). However, the two<br />

reports that test this hypothesis conflict. Minski et al. reported<br />

‘high food interest’ individuals (‘foodies’) rated quinine, sodium<br />

chloride and PROP as more intense than those with average or<br />

low food interest in a laboratory study (n=100), while Pickering<br />

et al. failed to find any difference in the bitterness of PROP<br />

impregnated discs sent via mail (n>900). These studies also<br />

differ in how high food affect individuals were identified: via a<br />

ratio of affective ratings <strong>for</strong> all foods to pleasant non-food item<br />

versus a difference score of mean liking <strong>for</strong> all foods minus mean<br />

liking <strong>for</strong> all non-foods. Here, we explore this question in 246<br />

subjects who completed a generalized hedonic survey (i.e. food<br />

& nonfood items) and whole mouth ratings <strong>for</strong> sucrose, quinine,<br />

and PROP. We characterized subjects as high affect using both<br />

approaches. Regardless of the categorization method, sucrose,<br />

quinine, and PROP intensity ratings did not differ by group in<br />

ANOVA. When groups were predicted in logistic regression<br />

(‘do higher taste ratings predict being a foodie?’) there was no<br />

evidence supporting this hypothesis. Adding the personality trait<br />

Sensation Seeking as a covariate did not alter these conclusions.<br />

We did find evidence of lower Sensation Seeking scores<br />

among foodies (as defined via the difference score), but further<br />

inspection suggests this was due to a small increase in the mean<br />

liking of pleasant non-food items when mean food liking was<br />

flat. These data fail to support the hypotheses that a) hypergeusic<br />

individuals show higher food related affect or that b) higher food<br />

affect predicts heightened taste response. Acknowledgements:<br />

funds from the Pennsylvania State University and NIH grant<br />

DC0010904.<br />

#P230 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

Effects of Food Neophobia on Salivary ph, Cortisol and<br />

Adrenal Level<br />

August Capiola, Bryan Raudenbush, Amanda Schultz<br />

Wheeling Jesuit University Wheeling, WV, USA<br />

Food neophobics (individuals reluctant to try novel foods)<br />

and food neophilics (individuals with an overt willingness<br />

to try novel foods) differ in several physiological aspects.<br />

Phenylthiocarbamide (PTC) tasting ability, a genetic<br />

predisposition, differs among the two groups with more food<br />

neophobics possessing this inherited trait. Food neophobics<br />

salivate less when presented with novel foods and have higher<br />

physiological stress responses to novel foods (increased pulse,<br />

GSR, and respirations). The present study assessed salivary<br />

pH, adrenal level and cortisol level in food neophobics, food<br />

neophilics and an average group, to determine whether such<br />

salivary flow and physiological stress reactions could partially be<br />

explained by such variables. Salivary mouth swab samples were<br />

obtained from 117 participants, who also completed the Food<br />

Neophobia Scale (FNS) to assess level of food neophobia. A<br />

significant MANCOVA result was found, F=2.47, p=.03. Further<br />

analysis revealed food neophobics had significantly higher levels<br />

of salivary cortisol compared to food neophilics and the average<br />

group, F(2,102)=7.53, p=.001. The finding that higher levels<br />

of the stress hormone cortisol are present in food neophobic’s<br />

saliva supports past research indicating a greater physiological<br />

stress reaction to novel food stimuli in these individuals. Future<br />

research should assess whether exposure to novel foods can<br />

decrease the level of salivary cortisol in food neophobics, as a<br />

way of promoting a more varied and healthful diet.<br />

#P231 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

The NIH Toolbox Brief Gustation Assessment Protocol<br />

Susan E. Coldwell 1 , Valerie B. Duffy 2 , Linda Bartoshuk 3 ,<br />

James W. Griffith 4 , Howard J. Hoffman 5<br />

1<br />

University of Washington School of Dentistry Seattle, WA, USA,<br />

2<br />

University of Connecticut College of Agriculture and Natural<br />

Resources Storrs, CT, USA, 3 University of Florida College of Dentistry<br />

Gainesville, FL, USA, 4 Northwestern University Feinberg School of<br />

Medicine Evanston, IL, USA, 5 National Institutes of Health, National<br />

Institute on Deafness and Other Communication Disorders Bethesda,<br />

MD, USA<br />

NIH Toolbox developed standardized, brief assessments <strong>for</strong><br />

sensory, motor, cognitive and emotional function that are<br />

designed <strong>for</strong> use in longitudinal studies, epidemiological<br />

research, and clinical trials. The sensory battery consists of<br />

brief assessments of gustation, olfaction, vision, audition, pain,<br />

and vestibular balance; all six can be administered within 30<br />

minutes, including 6 minutes <strong>for</strong> the assessment of gustation.<br />

The Gustation Assessment begins with instructions in use of the<br />

general Labeled Magnitude Scale by rating of remembered lights<br />

(dimly lit restaurant, well-lit room, brightest light ever seen).<br />

Four taste trials are then delivered: 1 mM Quinine HCl applied<br />

to the anterior tongue, 1 M NaCl applied to the anterior tongue,<br />

1 mM Quinine HCl whole mouth (sip and spit), and 1 M NaCl<br />

whole mouth. Rinsing with water is done between trials. As<br />

part of the NIH Toolbox national norming study, the Gustation<br />

Assessment was given to 1843 English-speakers and 240 Spanishspeakers.<br />

These included 494 subjects aged 12 to 15 years and<br />

509 aged 15 to 19 years. For 172 subjects, the battery was given<br />

twice to establish test-retest reliability. Preliminary intraclass<br />

correlations (ICC) indicate that the test is reliable <strong>for</strong> whole<br />

mouth ratings (ICC = 0.54 <strong>for</strong> Quinine and 0.57 <strong>for</strong> NaCl) and<br />

reliable <strong>for</strong> NaCl on the anterior tongue (ICC = 0.42). Quinine<br />

ratings <strong>for</strong> the anterior tongue were less reliable (ICC = 0.29),<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

119


ut this was always the first trial. For the entire norming sample,<br />

small, but statistically significant, declines in taste intensity<br />

with age were observed <strong>for</strong> each of the four trials (Pearson<br />

correlations from -0.1 to -0.2, p


were phenotyped <strong>for</strong> detection thresholds <strong>for</strong> sweet (sucrose),<br />

savory (monosodium glutamate), and salty (NaCl) taste using<br />

age-appropriate psychophysical testings, dietary habits, blood<br />

pressure and obesity. Preliminary analyses revealed that the<br />

detection threshold <strong>for</strong> sucrose, salt, and MSG are similar to<br />

that previously reported in adults and there were no differences<br />

observed between normal weight children and obese children in<br />

the detection thresholds <strong>for</strong> any of these basic tastes. In normalweight,<br />

but not obese children, salt detection thresholds were<br />

positively correlated with systolic blood pressure. As a group, the<br />

greater the waist circumference, the lower the sucrose detection<br />

threshold (the more sensitive the child was to sucrose) (p=0.02).<br />

No such relationships existed <strong>for</strong> salt or MSG in children.<br />

Whether the lower detection thresholds <strong>for</strong> sucrose are associated<br />

with stronger rein<strong>for</strong>cing value of sweet foods, as has been<br />

observed in adults, and understanding the link between salt taste<br />

thresholds and blood pressure, are important areas <strong>for</strong> future<br />

research. Acknowledgements: This project was funded by an<br />

investigator-initiated grant from Ajinomoto, Inc.<br />

#P235 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

Statistical Analysis of Factors Previously Described as<br />

Significant in the Ability to Taste Propylthiouracil Yields<br />

Roles <strong>for</strong> Age, Sex and Tas2r38 Haplotype, but not<br />

Fungi<strong>for</strong>m Papillae Density<br />

Nicole L. Garneau, Tiffany Derr<br />

Denver Museum of Nature & Science Denver, CO, USA<br />

The Genetics of Taste research study at the Denver Museum<br />

of Nature & Science is in a unique position to collect samples<br />

from a diverse population across a wide age range. Using this<br />

large population sample we set out to establish the scientific<br />

credibility of our community-based laboratory by replicating four<br />

previously reported statistically significant factors in the ability<br />

to taste propylthiouracil; 1)Age, 2)Sex, 3)Tas2r38 haplotype, and<br />

4)Fungi<strong>for</strong>m papillae density. Using regression analysis and the<br />

Student T-test we can replicate the role of age and sex in taste<br />

score (gLMS following a propylthiouracil taste test). Similarly,<br />

the presence of at least one dominant allele <strong>for</strong> the Tas2r38 gene<br />

is a significant predictor of taste. Finally, in order to decrease<br />

subjectivity, we developed a dichotomous key <strong>for</strong> objective<br />

analysis of fungi<strong>for</strong>m papillae density. Using this method we did<br />

not find that increased papillae density correlates to an increased<br />

propylthiouracil taste score. In conclusion, the ability to replicate<br />

the significance of age, sex and Tas2r38 haplotype, as well as the<br />

development of objective methodology <strong>for</strong> papillae analysis, all<br />

demonstrate the ability <strong>for</strong> citizen-scientists in a communitybased<br />

laboratory to collect, prepare and analyze data that can<br />

contribute to the field of chemoreception. In addition, we submit<br />

that using standardized methodology <strong>for</strong> fungi<strong>for</strong>m papillae<br />

density allows <strong>for</strong> a more objective analysis of morphological<br />

data. Using this methodology we find no relationship between<br />

fungi<strong>for</strong>m papillae density and propylthiouracil score in our<br />

data set. This data contradicts previously published studies and<br />

suggests that fungi<strong>for</strong>m papillae density may not be as reliable<br />

a metric <strong>for</strong> classifying taster status as previously thought.<br />

Acknowledgements: This work was supported by volunteer<br />

citizen-scientists at the Denver Museum of Nature & Science,<br />

and through support from 2008-2012 from R25 RR025066 NIH<br />

NCRR SEPA.<br />

#P236 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

The effects of temperature on sequential and mixture<br />

interactions between sucrose and saccharin<br />

Barry G Green 1,2 , Danielle Nachtigal 1<br />

1<br />

The John B. Pierce Laboratory New Haven, CT, USA,<br />

2<br />

Yale University School of Medicine New Haven, CT, USA<br />

The sweet taste of sucrose and saccharin has been shown to<br />

depend on stimulation of the T1R2-T1R3 receptor, but it is<br />

also clear that these two stimuli interact with the receptor in<br />

different ways. Most recently it was found that self-adaptation is<br />

temperature-dependent <strong>for</strong> sucrose but not <strong>for</strong> saccharin (Green<br />

& Nachtigal, 2012), and a previous study showed that high<br />

concentrations of saccharin can block the sweetness of sucrose<br />

(and of itself) and evoke a sweet water taste (Galindo-Cusperino<br />

et al, 2006). The aim of the present study was to determine if<br />

temperature modulates the ability of sucrose to cross-adapt<br />

saccharin and/or of saccharin to block the sweetness of sucrose<br />

and produce a sweet water taste. Subjects rated the sweetness<br />

and bitterness of 0.42 M sucrose, 3.2 mM saccharin, 100 mM<br />

saccharin, or binary mixtures of sucrose and the 2 concentrations<br />

of saccharin, with and without pre-exposure to themselves or<br />

each of the other stimuli. The variables of interest were the<br />

duration of pre-exposure (3 or 10 s) and solution temperature<br />

(37° or 21°C). The stimuli were sampled by dipping the tongue<br />

tip into the solutions, and intensity ratings were made on the<br />

gLMS be<strong>for</strong>e the tongue was retracted back into the mouth.<br />

The results confirmed the previous findings and showed that (1)<br />

the magnitude of sweet water taste (after exposure to 100 mM<br />

saccharin) is temperature-dependent, and (2) surprisingly, preexposure<br />

to sucrose <strong>for</strong> 3 or 10 sec appeared to counteract the<br />

ability of 100 mM saccharin to block sweetness, independent of<br />

temperature. These results support the hypothesis that sucrose<br />

and saccharin bind to at least 2 different sites on the T1R2-T1R3<br />

receptor and raise new questions about the factors that can<br />

affect excitatory and inhibitory interactions between these sites.<br />

Acknowledgements: Supported in part by NIH grant DC005002<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

121


#P237 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

Language Determines Whether Taste Sensitivity is Related<br />

to Moral Disgust<br />

Rachel S, Herz<br />

Department of Psychiatry and Human Behavior Providence, RI, USA<br />

The stimuli that trigger emotional disgust are currently debated.<br />

Most contested is whether responses to visceral triggers of<br />

disgust (e.g., cockroaches, disease) are fundamentally the same as<br />

responses elicited by moral transgressions (e.g., lying, cheating).<br />

To address this issue, the present study examined whether: (1)<br />

visceral and moral disgust share a common oral origin (the<br />

rejection of bitter tasting poisons); (2) verbal priming can alter<br />

whether disgust is experienced as a function of taste sensitivity.<br />

102 undergraduates completed a “Behavioral Situations<br />

Questionnaire” developed <strong>for</strong> the present research which<br />

compared “grossed out” and “angry” to assess three types of<br />

moral transgressions that varied in the degree to which a visceral<br />

disgust dimension was invoked (non-visceral, implied visceral,<br />

directly visceral), and two standard tests of disgust sensitivity.<br />

After completing the questionnaires, taste sensitivity was<br />

assessed with a bitter tasting compound (6-n-propylthiouracil;<br />

PROP). Results showed that the more bitter PROP tasted the<br />

more sensitive a participant was to visceral measures of disgust<br />

sensitivity, but not to moral disgust sensitivity. There were also<br />

no effects <strong>for</strong> taste sensitivity or type of moral transgression<br />

when “angry” was primed <strong>for</strong> evaluating the transgressions.<br />

However, when “grossed out” was primed, the more intense<br />

PROP tasted the more “grossed out” participants were by all<br />

transgressions, regardless of their visceral nature. This supports<br />

the proposition that moral and visceral disgust do not share<br />

a common oral origin, but shows that linguistic priming can<br />

trans<strong>for</strong>m a moral response into a viscerally repulsive event<br />

and that susceptibility to this priming varies as a function of<br />

an individual’s sensitivity to the origins of visceral disgust—<br />

bitter taste.<br />

#P238 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

Individual differences in the rate of salivary a-amylase<br />

production and its role in the perception of glucose polymers<br />

Trina J. Lapis, Michael H. Penner, Juyun Lim<br />

Oregon State University Corvallis, OR, USA<br />

Our recent data suggest that some humans can consistently<br />

taste glucose polymers, implying possible existence of a glucose<br />

polymer receptor. A potential confound of this idea is that the<br />

hydrolysis byproducts of the glucose polymers, i.e., glucose<br />

and/or maltose, may have, at least to some extent, influenced<br />

taste responsiveness of the tasters. The latter scenario is easily<br />

rationalized due to the catalytic activity of salivary a-amylase.<br />

The present study was designed to investigate (1) individual<br />

differences in the rate of a-amylase production and (2) the<br />

role that it may play in glucose polymer perception. Measured<br />

rate of salivary flow (mg/sec) and a-amylase activity per mg<br />

saliva were used to calculate the rate of a-amylase production<br />

(activity/sec). The same Ss rated the taste intensity of glucose,<br />

sucrose, and glucose polymer solutions (differing in average<br />

chain length) following a sip and spit procedure. Results showed<br />

large individual differences in the rate of a-amylase production<br />

(>30-fold). This can be attributed to the large differences in<br />

salivary flow rate (>18-fold) and a-amylase activity (>30-fold).<br />

Notably, average rates of a-amylase production were similar<br />

between the taster and non-taster groups. Further, within each<br />

group, responsiveness to glucose polymers did not appear to<br />

differ between individuals with high and low rates of a-amylase<br />

production. These findings suggest that salivary a-amylase plays<br />

an insignificant role in glucose polymer perception. Alternatively,<br />

it is possible that the chain lengths of the glucose polymers tested<br />

were too short, i.e., the effect of a-amylase on the hydrolysis was<br />

comparable between high and low a-amylase producers. This<br />

possibility is currently being explored in a follow-up experiment<br />

by using more complex carbohydrates as test stimuli.<br />

#P239 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

Evidence that humans can taste glucose polymers<br />

Juyun Lim, Trina J. Lapis<br />

Oregon State University Corvallis, OR, USA<br />

Previous findings of behavioral and electrophysiological studies<br />

have shown that rodents can taste solutions of Polycose and<br />

further can discriminate the taste of Polycose from that of<br />

sugars. Recent studies have also shown that the T1R2/T1R3<br />

single and double KO mice respond significantly to Polycose.<br />

Based on these data, the possible existence of a glucose polymer<br />

receptor has been proposed. In contrast, it has been assumed<br />

that glucose polymers are tasteless to humans, although it is<br />

known that they evoke a slight odor. During a preliminary study<br />

of the role of salivary a-amylase in the perception of glucose<br />

polymers, some Ss reported that the glucose polymers had<br />

“bread-” or “cereal-like” taste which they could differentiate<br />

from the sweet taste of sugars. The current study was there<strong>for</strong>e<br />

designed to measure individual differences in taste perception of<br />

various carbohydrates (glucose, sucrose, and glucose polymers<br />

of different chain length) and NaCl. Ss rated taste intensity<br />

of test solutions while their noses were clamped. The results<br />

showed that the perceived intensities of glucose, sucrose, and<br />

NaCl were significantly correlated (r=.68~.80, p


#P240 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

The role of gene expression in the human TAS2R38<br />

genotype-phenotype relationship<br />

Sarah V. Lipchock 1 , Andrew I. Spielman 2 , Julie A. Mennella 1 ,<br />

Danielle R. Reed 1<br />

1<br />

Monell Chemical Senses Center Philadelphia, PA, USA, 2 College of<br />

Dentistry, New York University New York, NY, USA<br />

Bitter taste is important <strong>for</strong> sensing and avoiding toxins, but the<br />

aversion to bitter can lead to low consumption of vegetables,<br />

an important source of phytochemicals. Single-nucleotide<br />

polymorphisms in the genes that code <strong>for</strong> the 25 human bitter<br />

receptors (TAS2Rs) result in different individual responses to<br />

taste stimuli. One example of this phenomenon is the TAS2R38<br />

gene, which encodes a receptor that recognizes compounds<br />

including 6-n-propylthiouracil (PROP) and goitrin, found<br />

in broccoli. Cell-based assays and psychophysical threshold<br />

measurements confirm that receptor with the amino acid<br />

sequence PAV responds to PROP, while the AVI <strong>for</strong>m does not.<br />

Heterozygous (PAV/AVI) individuals display a broad range of<br />

abilities to taste PROP. To examine the role of allele-specific<br />

gene expression in the TAS2R38 genotype-phenotype relationship<br />

we recruited healthy, non-smoking adults with the PAV/AVI<br />

diplotype (N=18). Psychophysical ratings <strong>for</strong> several bitter<br />

compounds and vegetable juices were related to levels of TAS2R<br />

gene expression from human fungi<strong>for</strong>m taste papillae. Increased<br />

levels of the PAV <strong>for</strong>m of TAS2R38 were positively correlated<br />

with the subject’s intensity ratings <strong>for</strong> PROP and broccoli juice.<br />

Expression of TAS2R43 was associated with ratings <strong>for</strong> caffeine,<br />

urea, denatonium benzoate and quinine. This relationship<br />

resulted from copy number variation of TAS2R43 and suggests<br />

that TAS2R43 acts as a proxy <strong>for</strong> bitterness sensitivity from the<br />

cluster of receptors on chromosome 12. Our data provide a link<br />

between genotype and phenotype in the taste system. Our work<br />

lays a foundation <strong>for</strong> future studies about regulatory mechanisms<br />

and implications of taste receptor expression as it relates to diet<br />

and vegetable consumption in humans. Acknowledgements:<br />

Supported by F32DC011975 (SVL), P30DC011735 (DRR) and<br />

R01DC011287 (JAM).<br />

#P241 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

Evidence that cooling affects the sweetness of sugars via 2<br />

separate mechanisms.<br />

Danielle J. Nachtigal 1 , Barry G. Green 1,2<br />

1<br />

The John B. Pierce Laboratory New Haven, CT, USA, 2 Department of<br />

Surgery, Yale University School of Medicine New Haven, CT, USA<br />

Previous experiments in our laboratory demonstrated that<br />

temperature can affect the sweetness of sucrose, glucose, and<br />

fructose solutions by modulating taste adaptation: Cooling from<br />

37°C to 21°C significantly increased the rate and amount of selfadaptation<br />

when the tongue tip was dipped into the solutions <strong>for</strong><br />

several seconds. However, cooling to 21°C did not reduce initial<br />

sweet taste intensity, and thus did not directly affect sensitivity<br />

to the sugars. In the present study we investigated whether<br />

cooling to colder temperatures would further accelerate sweet<br />

taste adaptation and/or begin to reduce sweet taste sensitivity.<br />

Subjects rated the sweetness of 0.42 M sucrose solutions at<br />

temperatures of 41°, 37°, 30°, 21°, 10° or 5°C after dipping<br />

the tongue tip <strong>for</strong> 0, 3, or 10 sec into either 0.42 M sucrose or<br />

pure dH 2<br />

O of the same temperature. Sweetness ratings were<br />

made prior to retracting the tongue back into the mouth. There<br />

were two main findings: (1) sweet taste adaptation increased<br />

significantly at 21°C (replicating our previous results), and<br />

(2) pre-exposure to pure dH 2<br />

O at 5° or 10°C significantly<br />

reduced sweet taste intensity independent of taste adaptation.<br />

These findings indicate that cooling interferes with the<br />

perception of sucrose sweetness in two different ways: mild<br />

cooling increases the rate of adaptation whereas more extreme<br />

cooling reduces the initial sensitivity to sucrose. Experiments<br />

are planned that will test the hypothesis that these two effects<br />

occur at different stages of the transduction cascade, with the<br />

<strong>for</strong>mer affecting the binding of sucrose to the transmembrane<br />

region of the T1R2-T1R3 receptor, and the latter modulating a<br />

temperature-sensitive intracellular mechanism that is common to<br />

G-protein coupled receptors (e.g., TRPM5). Acknowledgements:<br />

NIH grant RO1 DC005002<br />

#P242 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

The Chemosensory Component in the 2012 National Health<br />

and Nutrition Examination Survey (NHANES): Test-Retest<br />

Analysis and Comparison with Laboratory-Based Measures<br />

Shristi Rawal 1 , Howard J Hoffman 2 , Kathy Bainbridge 2 ,<br />

Valerie B Duffy 1<br />

1<br />

University of Connecticut/Allied Health <strong>Sciences</strong> Storrs, CT, USA,<br />

2<br />

National Institute on Deafness and Other Communication Disorders/<br />

Division of Scientific Programs Bethesda, MD, USA<br />

The NHANES now includes a standardized chemosensory<br />

protocol with brief assessments of taste intensity and odor<br />

identification conducted in a mobile exam center. After<br />

orientation to the general Labeled Magnitude Scale and scaling<br />

intensity of LED-generated lights, participants report intensities<br />

of 1 M NaCl and 1 mM quinine hydrochloride applied to the<br />

tongue tip and these plus 0.32 M NaCl with the whole mouth.<br />

Olfactory dysfunction is assessed with two 4-item, scratchand-sniff<br />

tests (Pocket Tests (PT), Sensonics, Inc) with food<br />

(strawberry, chocolate, onion, grape) and common (leather,<br />

soap, smoke, natural gas) odors. Here we examined test-retest<br />

reliabilities of the chemosensory protocol, compared the PT<br />

with an olfactometer identification task, and added scaling<br />

to all odor tests. Fifty adults (mean age=27, range 18-62 yrs),<br />

primarily Caucasian and Asian, college-educated and reportedly<br />

healthy, were tested in a laboratory twice within two weeks.<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

123


The taste test intraclass correlations (ICC, single measures)<br />

ranged from moderate to substantial agreement (0.47 NaCl,<br />

0.50 quinine tongue tip; 0.75 quinine whole mouth). Both<br />

PT trials classified 98% of the participants as normosmic.<br />

There were good correlations between the individual PT odor<br />

intensities (ranging 0.44 <strong>for</strong> grape to 0.72 <strong>for</strong> gas) and a moderate<br />

overall ICC (0.56). The PT and olfactometer odor intensities<br />

were correlated, averaging 0.5 (ranging 0.28 <strong>for</strong> chocolate to<br />

0.65 <strong>for</strong> grape). Correct and incorrect odor identification was<br />

consistent across the PT and olfactometer tests, averaging 86%<br />

agreement. These findings show that, in ideal testing situations,<br />

the NHANES protocol had very good reliability and the Pocket<br />

Test corresponded reasonably well among normosmics with a<br />

measure having more stimulus control. Acknowledgements:<br />

NIDCD/NIH<br />

#P243 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

Efficacy of sodium and glutamate in reducing bitterness in<br />

children and adults<br />

Kristi M. Roberts, Phoebe S. Mathew, Corrine J. Mansfield ,<br />

Danielle R. Reed , Julie A. Mennella<br />

Monell Chemical Senses Center Philadelphia, PA, USA<br />

A central challenge of administering medicine to children is a<br />

‘matter of taste’ because drugs, by their very nature, often taste<br />

unpleasant with bitter taste being the primary culprit. As part<br />

of an ongoing study on individual differences in bitter taste<br />

perception, we present here preliminary data on ratings of the<br />

bitterness of urea and propylthiouracil (PROP) with and without<br />

the addition of a sodium gluconate or glutamate (putative<br />

blockers) in a racially diverse group of 3- to 10-year old children<br />

and adults. Using <strong>for</strong>ced-choice procedures, each child and adult<br />

was presented with all possible pairs of the four solutions <strong>for</strong><br />

each bitter agent (e.g., 0.5M urea, 0.3M sodium gluconate, 0.5<br />

M urea+0.3M sodium gluconate, and water), one pair at a time,<br />

and asked to indicate which of the pair tasted more bitter. The<br />

data <strong>for</strong> each bitter stimulus were expressed as the proportion of<br />

children or mothers that chose one member of the pair as tasting<br />

more bitter and from this, each of the four solutions were ranked<br />

according to subject’s ratings (1=least bitter; 4=most bitter).<br />

Adults also used gLMS to rate taste qualities of each solution<br />

in another session. Preliminary analysis revealed that most<br />

children (75%) were able to complete the task; the vast majority<br />

of those who did not complete the task were of the younger age<br />

(


#P245 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

#P246 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

Glutamate Detection Thresholds Are Altered by the<br />

Addition of 5’-Ribonucleotides<br />

Ashley A Sharples 1 , Suzie Alarcon 1 , Paul AS Breslin 1,2<br />

1<br />

Rutgers University New Brunswick, NJ, USA, 2 Monell Chemical<br />

Senses Center Philadelphia, PA, USA<br />

Umami taste intensity of glutamate is synergistically increased<br />

in humans by the addition of inosine 5’-monophosphate<br />

(IMP) disodium salt and guanosine 5’-monophosphate (GMP)<br />

disodium salt, two purinergic ribonucleotides. We hypothesized<br />

that the addition of these and other ribonucleotides would<br />

decrease the absolute detection threshold of L-glutamic acid<br />

potassium salt (MPG) when in admixture; thus, sensitivity to<br />

glutamate would be increased. We first measured the absolute<br />

threshold to MPG and compared this to the threshold <strong>for</strong> MPG<br />

in the presence of a constant background level of ribonucleotide<br />

(3 mM). All thresholds were measured twice in all subjects.<br />

We found the additions of inosine monophosphate<br />

(IMP), guanosine monophosphate (GMP), and adenosine<br />

monophosphate (AMP), lowered the MPG threshold <strong>for</strong> every<br />

subject and were statistically significant p <br />

GMP> AMP >> UMP // CMP. We will explore associations<br />

among perceptual differences in synergistic effects of various<br />

ribonucleotides with genetic variations in individual oral<br />

glutamate receptors. Acknowledgements: Funded in part by<br />

NIH DC 02995 to PASB.<br />

Transient top-down modulation of gustatory sensitivity<br />

in humans<br />

Maria G Veldhuizen 1,2 , Elias D Lustbader 1 , Dana M Small 1,2,3<br />

1<br />

THE JOHN B PIERCE LABORATORY NEW HAVEN, CT, USA,<br />

2<br />

Department of Psychiatry, Yale School of Medicine NEW HAVEN,<br />

CT, USA, 3 Department of Psychology, Yale University NEW HAVEN,<br />

CT, USA<br />

Perceptual systems are dynamic. Sensitivity may decrease due to<br />

adaptation or increase as a result of directed attention. In three<br />

studies we tested the hypothesis that sustained directed attention<br />

to oral stimulation would influence gustatory sensitivity. In<br />

study one, six subjects rated the intensity of taste stimuli<br />

(sucrose, sucralose, citric acid, sodium chloride, monosodium<br />

glutamate) three times be<strong>for</strong>e and after they per<strong>for</strong>med a triangle<br />

task requiring sustained directed attention to oral sensation. In<br />

the triangle test subjects were presented with two qualitatively<br />

similar flavor stimuli and asked to “pick the odd one out”. They<br />

repeated this task 12 times and then rated the intensity of the<br />

taste stimuli again. Subjects returned to the lab four times and<br />

repeated this procedure. As predicted, taste stimuli were rated<br />

as more intense post- vs. pre- triangle test (p=.009). However,<br />

the effect was transient in that taste stimuli were rated as less<br />

intense during the pre-triangle test vs. the previous days’ posttriangle<br />

test (p=.000). In study two we replicated these findings<br />

with a sample of 9 subjects (p=.000 and p=.037). In study three<br />

we tested whether consumption of the flavor stimuli without<br />

the per<strong>for</strong>mance of the triangle test would result in sensitivity<br />

changes. Consistent with the hypothesis that it is the per<strong>for</strong>mance<br />

of the discrimination task, rather than mere exposure to flavors<br />

that produces the increased sensitivity to taste, no significant<br />

increases in taste intensity were observed post- vs. pre- exposure<br />

(p = .115). These results demonstrate that per<strong>for</strong>ming a difficult<br />

chemosensory task results in a transient increase in taste<br />

sensitivity, which we suggest reflects a temporary effect of topdown<br />

modulation on increasing the efficiency of central taste<br />

processing. Acknowledgements: Supported by NIDCD grant<br />

R01 DC006706<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

125


#P247 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

#P248 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

The Role of Glutamate in Infant Satiation: a Model System<br />

Alison K. Ventura 1, 2 , Loma B. Inamdar 2 , Julie A. Mennella 2<br />

1<br />

Monell Chemical Senses Center Philadelphia, PA, USA,<br />

2<br />

Drexel University Philadelphia, PA, USA<br />

We recently discovered that human infants consumed less<br />

to satiation when feeding <strong>for</strong>mulas high in free glutamate<br />

(extensively protein hydrolysate <strong>for</strong>mulas; ePHF) than when<br />

consuming a <strong>for</strong>mula low in free glutamate, such as standard<br />

cows’ milk <strong>for</strong>mulas (CMF). The purpose of this study was to<br />

use this model system to characterize the timing and patterning<br />

of cues infants use to signal satiation. In this within-subjects<br />

study, 41 infants


#P249 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

#P250 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

Re-Engineering of Olfactory Receptor OlfCc1 Toward<br />

Directed Ligand Selectivity<br />

Allison P Berke 1 , Francine Acher 2 , Hugues O. Bertrand 3 , John Ngai 4<br />

1<br />

Joint Graduate Group in Bioengineering, University of Cali<strong>for</strong>nia<br />

Berkeley, CA, USA, 2 Laboratoire de Chimie et Biochimie<br />

Pharmacologiques et Toxicologiques, Unite Mixte de Recherche<br />

8601, Centre National de la Recherche Scientifique, Universite Rene<br />

Descartes-Paris V Paris, France, 3 Accelrys Orsay, France, 4 Department<br />

of Molecular and Cell Biology and Helen Wills Neuroscience Institute<br />

Berkeley, CA, USA<br />

Fish sense food cues in their aqueous environment using family<br />

C GPCR olfactory receptors. These C family receptors are<br />

characterized by a large N-terminal “Venus flytrap” domain.<br />

Zebrafish olfactory receptor OlfCc1 is a broadly-expressed<br />

ortholog of mammalian V2R2, which shares sequence similarity<br />

with the human Calcium-sensing Receptor (CaSR). C family<br />

receptors are predicted to respond to amino acids, as do the<br />

previously identified OlfCa1 and CaSR. The expression of<br />

OlfCc1 in the entire microvillous olfactory neuron population<br />

in the zebrafish, as well as its sequence homology with CaSR,<br />

make it an interesting target <strong>for</strong> de-orphaning and engineering,<br />

as it could play a generalized behavioral role in zebrafish<br />

chemosensation. In silico modeling of OlfCc1 identified the<br />

receptor’s binding pocket and residues likely to be directly<br />

involved in ligand binding. OlfCc1 was then cloned into a CMVI<br />

FLAG-tagged expression vector and expressed in HEK293<br />

cells. Calcium imaging per<strong>for</strong>med on these cells using Fluo-4<br />

calcium-sensitive dye revealed the calcium-dependent binding<br />

profile of OlfCc1, which includes amino acids. Amino acid<br />

point mutations were then introduced to OlfCc1, with the<br />

aim of broadening the receptor’s binding specificity. These<br />

mutations succeeded in altering the sensitivity and specificity<br />

of OlfCc1 in accordance with predictions. In conclusion, the<br />

binding specificity and sensitivity of OlfCc1 can be selectively<br />

engineered. Additionally, the combination of in silico homology<br />

modeling and calcium imaging that constitute this method can<br />

be applied to other C family GPCRs, to directly engineer ligand<br />

binding capability. Acknowledgements: NSF GRFP and the NIH<br />

Interactions at the olfactory receptor level contribute to the<br />

coding of odorant mixtures<br />

fouzia El Mountassir 1 , christine Belloir 1 , loic Briand 1 ,<br />

thierry Thomas Danguin 1 , anne marie Le BON 1<br />

1<br />

Centre des <strong>Sciences</strong> du Gout et de l’Alimentation DIJON, France, 3<br />

Numerous studies reported that the perceptual characteristics<br />

of odorant mixtures are often different from those of their<br />

individual compounds; e.g. the mixture intensity can be higher<br />

or lower than the arithmetic sum of each component’s intensity.<br />

These findings raise the question how odorants in mixtures are<br />

detected and encoded at the peripheral level of the olfactory<br />

system. We investigated this question through the measurement<br />

of human olfactory receptor (OR) responses to two specific<br />

binary mixtures of aldehydes: (i) octanal and citronellal, known<br />

to induce a configural perception in rats (Kay et al., 2003) and<br />

masking effects in humans (Burseg et al., 2009); (ii) octanal<br />

and methional, known to induce masking effects in humans<br />

(Burseg et al., 2009). We used a heterologous expression system<br />

(HEK293T cells) in which OR (OR1G1, OR52D1, OR2W1<br />

and OR1A1) were transfected transiently. Responses of OR to<br />

odorants applied alone or in mixtures were measured by calcium<br />

imaging. The results showed various interactions at the OR<br />

level. When octanal was mixed with citronellal, the OR response<br />

intensity was reduced thus showing subtraction, compromise or<br />

partial addition, depending on the OR and the concentrations of<br />

odorants. Interestingly, the mixture of octanal and methional was<br />

found to induce mostly synergy, whatever the OR. These data<br />

strengthen the hypothesis that interactions can occur at the OR<br />

level and could there<strong>for</strong>e contribute significantly to the olfactory<br />

coding of odorant mixtures. Acknowledgements: This work is<br />

funded by the National Institute of Agricultural Research and<br />

the region of Burgundy<br />

#P251 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

Class-specific regulation of a zebrafish olfactory receptor gene<br />

Stefan H. Fuss, Xalid Bayramli, Nuray Sögünmez<br />

Bogazici University, Molecular Biology and Genetics Istanbul, Turkey<br />

Olfactory sensory neurons (OSNs) typically express a single<br />

allele of a single olfactory receptor (OR) gene from a much larger<br />

genomic repertoire; a phenomenon that is not well understood.<br />

Experimental evidence suggests that OR expression is controlled<br />

by a combination of long- and short-range regulatory sequences.<br />

We used promoter bashing in transgenic zebrafish to identify<br />

proximal regulatory sites within the OR101-1 gene promoter.<br />

Positive regulatory sites are located within the first 500 bp<br />

upstream of the transcription start site, while more distant<br />

sequences confer a repressive effect on transgene expression, even<br />

in the presence of strong genomic enhancers. Interestingly, the<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

127


epressive sequence contains binding sites <strong>for</strong> the transcriptional<br />

repressor zbtb7b, which is abundantly expressed in the peripheral<br />

zebrafish olfactory system. Preliminary results suggest that<br />

negative regulation confers a higher specificity of transgene<br />

expression to ciliated OSNs. A major distinction within the OR<br />

repertoire can be made between evolutionary ancient class I ORs<br />

and the group of class II receptors, which massively expanded in<br />

the tetrapod lineage. OR101-1 appears to be the only class II OR<br />

in zebrafish. OSNs that fail to express a functional OR usually<br />

undergo a second OR gene choice resulting in heterogeneity of<br />

OR expression and widespread axonal projection of transgenic<br />

OSN axons to olfactory bulb glomeruli. Transgenic OSNs<br />

expressing a deletion allele of the OR101-1 gene in contrast<br />

target a single specific glomerulus, in the zebrafish olfactory<br />

bulb suggesting that second OR gene choice is severely restricted<br />

in those OSNs. The results imply that distinct regulatory<br />

mechanisms evolved concomitantly with the appearance of<br />

the first class II OR in the teleost lineage. Acknowledgements:<br />

TÜBITAK, The Scientific and Technological Research Council<br />

of Turkey, Grant Awards: 107T760, 112T168<br />

#P252 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

Olfactory Receptor (OR) switching is influenced by genome<br />

position in olfactory-placode (OP)-derived cells.<br />

Robert P. Lane, Seda Kilinc<br />

Wesleyan University Middletown, CT, USA<br />

We previously characterized olfactory receptor (OR) expression<br />

in the OP6 and OP27 cell lines and made two general<br />

observations: OR choice is not a heritable property; and the<br />

range of OR representation in OP populations appears biased<br />

<strong>for</strong> a subset of the full OR repertoire. We used custom arrays<br />

and deep sequencing to analyze the complete expressed OR<br />

repertoire in OP cultures. OP6 and OP27 cell lines have<br />

significant overlap in OR representation, consistent with similar<br />

pre-specification of the two founder cells, which had been<br />

isolated from a common developmental milieu. However, OR<br />

representation in OP cultures is not constrained by presumptive<br />

expression zones within mouse olfactory epithelium, as<br />

might be predicted if the range of OR choices had been prespecified<br />

by developmental (i.e., spatial) cues. Instead, we find<br />

strong evidence <strong>for</strong> ‘position-effects’: neighboring OR genes<br />

are significantly over-represented in divergent populations,<br />

suggesting a tendency to switch within versus across OR<br />

clusters. Surprisingly, we do not observe differences in common<br />

epigenetic marks between “active” versus “inactive” ORs, nor<br />

does locus positioning relative to nuclear chromocenters appear<br />

to be predictive of selection probability. We suggest a switching<br />

model in which the epigenetic microenvironment established by<br />

previous OR transcription increases the likelihood of selection/<br />

re-selection within that genome region. We hypothesize that<br />

the OP6 and OP27 founder cells had been similarly specified,<br />

initially leading to a similar bias <strong>for</strong> OR switching, but that in<br />

the absence of further developmental cues during subsequent<br />

culturing, probabilistic switching histories have resulted in<br />

slow evolution from initial biases, thus independently evolving<br />

OR subrepertoires. Acknowledgements: NIH R01-DC006267<br />

NSF 0842868<br />

#P253 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

Blockade of Insect Odorant Receptor Currents by<br />

Amiloride Derivatives<br />

Gregory M Pask 1 , Yuriy V Bobkov 2 , Elizabeth A Corey 2 ,<br />

Barry W Ache 2,3 , Laurence J Zwiebel 1,4<br />

1<br />

Department of Biological <strong>Sciences</strong>, Vanderbilt University Nashville,<br />

TN, USA, 2 Whitney Laboratory, Center <strong>for</strong> Smell and Taste, and<br />

McKnight Brain Institute, University of Florida Gainesville, FL, USA,<br />

3<br />

Departments of Biology and Neuroscience, University of Florida<br />

Gainesville, FL, USA, 4 Department of Pharmacology, Vanderbilt Brain<br />

Institute and Center <strong>for</strong> Human Genetics, Institutes of Chemical Biology<br />

and Global Health and Program in Developmental Biology, Vanderbilt<br />

University Medical Center Nashville, TN, USA<br />

Insect odorant receptors (ORs) function as heteromeric odorantgated<br />

ion channels consisting of a conserved coreceptor, Orco,<br />

and an odorant-sensitive tuning subunit. Although some<br />

OR modulators have been identified, an extended library of<br />

pharmacological tools is currently lacking and would aid in<br />

furthering our understanding of insect OR complexes. Using<br />

whole-cell patch clamp techniques, we demonstrate that<br />

amiloride and several derivatives, which have been extensively<br />

used as blockers <strong>for</strong> various ion channels and transporters, also<br />

block odorant-gated currents from two OR complexes from<br />

the malaria vector mosquito Anopheles gambiae, AgOr48 +<br />

AgOrco and AgOr65 + AgOrco. In addition, currents from<br />

both heteromeric OR complexes were susceptible to HMA<br />

blockade when activated by VUAA1, an agonist that targets<br />

the Orco channel subunit. HMA was also capable of blocking<br />

VUAA1-evoked currents of Orco homomers from 4 different<br />

insect orders, demonstrating that HMA blockade is not unique<br />

to AgOR complexes. Additionally, amiloride derivatives have<br />

provided insights into the properties of both the channel pore<br />

and spontaneous gating across different insect OR complexes.<br />

Amiloride derivatives there<strong>for</strong>e represent a valuable class of<br />

channel blockers that can be used to further investigate the<br />

pharmacological and biophysical properties of insect OR<br />

function. Acknowledgements: This work was supported by the<br />

National Institute of Allergy and Infectious Disease [AI056402]<br />

to L.J.Z and the National Institute on Deafness and Other<br />

Communication Disorders (NIDCD) [DC001655] to B.W.A. at<br />

the National Institutes of Health, and the Foundation <strong>for</strong> the<br />

National Institutes of Health through the Grand Challenges<br />

in Global Health Initiative [VCTR121] to L.J.Z. G.M.P was<br />

supported by the NIDCD through an NRSA F31 [DC011989].<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

128


#P254 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

Functional Analysis Of Nematode GPCRS In Yeast<br />

Muhammad Tehseen, Alisha Anderson, Mira Dumancic,<br />

Lyndall Briggs, Stephen Trowell<br />

CSIRO Food Futures Flagship and Division of Ecosystem <strong>Sciences</strong>,<br />

Black Mountain Laboratories Canberra, Australia<br />

The yeast Saccharomyces cerevisiae has been used extensively <strong>for</strong><br />

ligand screening of human G-protein coupled receptors, due to<br />

its ease of genetic manipulation, low cost, rapid growth, and<br />

eukaryotic secretory pathway. Although the Caenorhabditis elegans<br />

genome was sequenced 13 years ago and encodes over 1,000<br />

GPCRs, of which several hundred are believed to respond to<br />

volatile organic ligands, only one of these receptors, ODR-10,<br />

has been linked to a volatile ligand, 2,3-butanedione. ODR-3 is a<br />

G-protein a subunit believed to be involved in odorant detection<br />

and activated by ODR-10. Here we report the functional<br />

coupling of ODR-10 to the yeast pheromone signalling pathway<br />

using the yeast - C. elegans chimaeric Ga subunit (GPA-<br />

1:ODR-3). Interactions between ODR-10, ODR-3 and the<br />

chimaera were confirmed using the split ubiquitin yeast twohybrid<br />

system. We also report the tailoring of a Saccharomyces<br />

cerevisiae strain <strong>for</strong> the analysis of C. elegans chemoreceptor<br />

function. In this study, a yeast gpa1D ste2D sst2D far1D quadruple<br />

mutant strain was constructed to efficiently couple nematode<br />

olfactory receptors with the yeast signalling pathway. We<br />

used two different reporters: green fluorescent protein and<br />

ß-galactosidase, to verify activation of the signal transduction<br />

pathway by ligand activated GPCR interactions. With this<br />

heterologously engineered yeast system, we aim to accelerate the<br />

de-orphaning of C. elegans GPCR proteins.<br />

#P255 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

Differentiating activation of intracellular signaling pathways<br />

using calcium dynamics<br />

Kirill Ukhanov 1 , Yuri Bobkov 1 , Elizabeth A. Corey 1 , Barry W. Ache 1,2<br />

1<br />

University of Florida, Center <strong>for</strong> Smell and Taste Gainesville, FL,<br />

USA, 2 University of Florida, Depts. of Biology and Neuroscience<br />

Gainesville, FL, USA<br />

Mammalian olfactory receptors (ORs) appear to have the<br />

capacity to couple to multiple G protein-coupled signaling<br />

pathways, more specifically phosphoinositide-dependent<br />

signaling in addition to canonical cAMP-dependent signaling,<br />

in a ligand-selective manner. To better understand the<br />

mechanisms and molecular range of such ligand selectivity, we<br />

developed a heterologous expression system with differential<br />

readout of intracellular calcium changes. We expressed the<br />

mouse eugenol receptor (mOREG) in HEK293T cells together<br />

with Ga15 [phospholipase-C (PLC) pathway] and/or Gaolf<br />

[adenylate cyclase (AC) pathway], leading to intracellular<br />

calcium release or calcium influx through a cyclic nucleotidegated<br />

channel mutant deficient in Ca-CaM negative feedback,<br />

respectively. Eleven known mOREG agonists were tested,<br />

including eugenol, its analogs, and structurally dissimilar<br />

compounds (mousse cristal, nootkatone, orivone). PLCdependent<br />

responses differed dynamically [e.g., eugenol<br />

(t rise<br />

= 3.55 ± 0.13 sec; t decay<br />

= 72.42 ± 19.01 sec) from ACdependent<br />

responses (t rise<br />

= 90.83 ± 9.67 sec; t decay<br />

= 167.15 ±<br />

6.18 sec)], allowing them to be distinguished when Ga15 and<br />

Gaolf were co-expressed. This difference persisted across ligand<br />

concentration. All agonists tested activated both pathways [e.g.,<br />

EC 50<br />

, eugenol: 76 ± 12 µM (PLC), 78 ± 26 µM (AC)], showing<br />

that mOREG can couple to different G proteins expressed in the<br />

same cell. A larger scale screening is under way to identify and<br />

characterize potential OR-ligand combinations that differentially<br />

activate downstream signaling pathways. Acknowledgements:<br />

NIDCD DC001655, DC005995<br />

#P256 POSTER SESSION V:<br />

HUMAN TASTE PSYCHOPHYSICS;<br />

OLFACTION RECEPTORS; TASTE DEVELOPMENT<br />

Profiling of OR gene expression in the human<br />

olfactory epithelium<br />

Françoise Wilkin 1 , Christophe Verbeurgt 2 , Pierre Chatelain 1<br />

1<br />

ChemCom S.A. Brussels, Belgium, 2 Hôpital Erasme Brussels, Belgium<br />

Background: Olfactory recognition is mediated by a<br />

large repertoire of olfactory receptors (ORs). The human<br />

genome contains 851 OR loci. More than 50% of the loci<br />

are annotated as nonfunctional due to frame-disrupting<br />

mutations. Furthermore some missense haplotypic alleles can<br />

be nonfunctional due to a substitution of key amino acids<br />

governing protein folding or interaction with signal transduction<br />

components. Beyond their role in odor recognition, functional<br />

ORs are also required <strong>for</strong> a proper targeting of olfactory neuron<br />

axons to their corresponding glomeruli in the olfactory bulb<br />

(Feinstein et al, 2004). There<strong>for</strong>e, profiling of OR gene expression<br />

in the olfactory epithelium provides an opportunity to select<br />

frequently expressed and potentially functional ORs <strong>for</strong> large<br />

deorphanization campaign. Methods: An AB TaqMan® Low<br />

Density Array (TLDA) containing probes <strong>for</strong> 356 predicted OR<br />

loci was designed to investigate the chemosensory receptor gene<br />

expression in olfactory epithelium tissues from 8 individuals.<br />

Total RNA isolation, DNase treatment, RNA integrity<br />

evaluation and reverse transcription in cDNA were per<strong>for</strong>med<br />

<strong>for</strong> these 8 samples. Then 384 gene targets (including reference<br />

genes <strong>for</strong> normalization) were analysed using the same RT-qPCR<br />

plat<strong>for</strong>m. Results: The expression of 200 (56%) human OR<br />

genes was observed in these olfactory epithelia, among which<br />

114 were robustly expressed in all tested individuals. No relation<br />

between OR gene expression and age or sex was observed. Most<br />

of the ORs (>80%) deorphanised at Chemcom or described in<br />

the literature were found in the expressed set.<br />

POSTER PRESENTATIONS<br />

<strong>Abstracts</strong> are printed as submitted by the author(s).<br />

129


Author Index<br />

Abdelhamid, Mostafa – 38<br />

Abdus-Saboor, Ishmail – P39<br />

Abe, Keiko – P33<br />

Abraham, Michael H – P129<br />

Abrol, Ravinder – 4<br />

Ache, Barry W. – P68, P253, P255<br />

Acher, Francine – P249<br />

Ackels, Tobias – P191, P193<br />

Ahn, Christina – 45<br />

Åhs, Fredrik – P144<br />

Aiello, Salvatore – P1, P2<br />

Alarcón, Laura K. – P180<br />

Alarcon, Suzie – P226, P171, P245<br />

Albeanu, Dinu – 34, 36<br />

Albertowski, Katja – P19<br />

Alcaraz-Zubeldía, Mireya – P18<br />

Algarra, Laure – P175<br />

Allahverdizadeh, Masoud – P89<br />

Allen, Alissa L – P30, P227, P229<br />

Amano, Ryohei – P70, P107<br />

Anderson, Alisha – P254<br />

Anderson, Catherine B. – P165<br />

Anholt, Robert R H – P248<br />

Antenucci, Rachel G – P228<br />

Arshamian, Artin – P66<br />

Ashby, Sarah E – P86, P101<br />

Asson-Batres, Mary Ann – P55<br />

Atallah, Elias – P93<br />

Auclair, François – P93<br />

Avrillier, Julie – P175<br />

Ayabe-Kanamura, Saho – P142<br />

Bachmanov, Alexander A. – P33, P163<br />

Bacigalupo, Juan – P179, P195<br />

Bainbridge, Kathy – P242<br />

Bains, Gurprit – P1<br />

Bajayo, Alon – 27<br />

Baker, Harriet – P16<br />

Bales, Michelle B. – P207<br />

Barham, Henry P – P115<br />

Barkai, Edi – P113<br />

Barlow, Linda – 27, 29, 45, P225<br />

Barnes, Dylan C – P87<br />

Bartoshuk, Linda M – P12, P13, P25, P231<br />

Baumgart, Sabrina – P44, P49<br />

Bayramli, Xalid – P251<br />

Bazhenov, Maxim – P108<br />

Beauchamp, Gary K. – 57, P163<br />

Beauchamp, Jonathan – P133, P139<br />

Beers, Nicole K – P127<br />

Behrens, Maik – P159<br />

Belloir, Christine – P161, P250<br />

Belluscio, Leonardo – 51<br />

Ben-Shahar, Yehuda – 54, 55<br />

Bendahmane, Mounir – P88<br />

Benedict, Christian – P128<br />

Bennegger, Willi – P69<br />

Béno, Noëlle – P63<br />

Bensafi, Moustafa – P63<br />

Berke, Allison P – P249<br />

Bertrand, Hugues O. – P249<br />

Beynon, Robert J – 18, P192<br />

Bharti, Kapil – P71<br />

Biggs, Lindsey M – P196<br />

Bills, Meghan L – P125<br />

Bintig, Willem – P44<br />

Birgersson, Göran – P187<br />

Birnbaumer, Lutz – 31<br />

Bizon, Jennifer L – P209<br />

Blumrich, Anna – P56<br />

Bobkov, Yuriy V – P68, P253, P255,<br />

Boehm, Ulrich – P159<br />

Boesveldt, Sanne – P38, P181<br />

Bond, Amanda E. – P166<br />

Bone, Richard – P5<br />

Bower, Emily S. – P7<br />

Boyles, Diana – P232<br />

Bozza, Thomas – 28, P89<br />

Bradley, Robert M. – P77, P84, P214, P216<br />

Brand, Mark H – P34<br />

Brasser, Susan M. – P74<br />

Braubach, Oliver R. – P89<br />

Breslin, Paul AS – P171, P173, P175,<br />

P245, P226<br />

Breza, Joseph M – 46<br />

Briand, Loïc – P161, P250<br />

Briggs, Lyndall – P254<br />

Brignall, Alexandra – P194<br />

Brown, Marsalis – P52<br />

Brunert, Daniela – 30, P90<br />

Brünner, Yvonne F. – P128<br />

Bryant, Bruce – P124<br />

Buettner, Andrea – P133, P139<br />

130


Author Index, continued<br />

Burke, Allison – P134<br />

Byrd-Jacobs, Christine A – P104, P111, P190<br />

Byrnes, Nadia K – P30, P229<br />

Cai, Huan – 47<br />

Cain, William S – P129<br />

Callaham, Amy E – P127<br />

Cameron, E. Leslie – P15<br />

Capiola, August – P8, P141, P230<br />

Carson, Ellen – P21<br />

Casey, Elizabeth M. – P199<br />

Castillo, David – 45<br />

Castro, Jason B – P130<br />

Castro, Norma – P74<br />

Cave, John W – P16<br />

Cazakoff, Brittany N – 35<br />

Cervantez, Melissa – P17<br />

Chae, Hong goo – 36<br />

Chai, Jinghua – P152, P200<br />

Chamero, Pablo – 31<br />

Chang, Jen – P50<br />

Chatelain, Pierre – P256<br />

Chaudhari, Nirupa – 22, 25, P151<br />

Chen, Chien-Fu F – P91<br />

Chen, Denise – P26, P27<br />

Chen, Jennifer – P26, P27<br />

Chen, Kepu – P131<br />

Chen, Zhixiong – 46<br />

Chennubhotla, Chakra S – P130<br />

Cheong M, MJ – P126<br />

Chirdon, Patrick – 47<br />

Chiu, Hong-Yan – P102<br />

Chokshi, Varun – P51<br />

Christensen, Michael – P123<br />

Churakov, Gennady A. – P183<br />

Cichy, Annika – P191, P193<br />

Ciesinski, Alexa – P119<br />

Claeson, Anna-Sara – P114<br />

Clark, Adam A. – P167<br />

Cleland, Thomas A. – P32, P92<br />

Clepce, Marion – P20<br />

Cloutier, Jean-François – P194<br />

Cobb, M – P41<br />

Cohen, Lawrence B. – P89, P109<br />

Cohen, Noam A. – 57<br />

Cohen, Shereen J. – P15<br />

Cohen, Yaniv – P113<br />

Coldwell, Susan E. – P231<br />

Cole, Sydni M. – P65<br />

Collins, David – P215<br />

Cometto-Muñiz, Jorge E – P129<br />

Cong, Wei-na – 47<br />

Connelly, Timothy – P40, P47<br />

Contreras, Robert J. – P207<br />

Cooper, Sarah E – P115<br />

Corey, Elizabeth A – P253, P255<br />

Corson, James A. – P77, P214<br />

Corson, Sara L – P216<br />

Coureaud, Gérard – P63, P201, P205<br />

Cowart, Beverly J. – P132<br />

Craven, Deirdre – P119<br />

Craw<strong>for</strong>d, Ryan – P132<br />

Croy, Ilona – P97<br />

Cruickshanks, Karen J. – P14, P184<br />

Crump, Kerensa L – 35<br />

Cut<strong>for</strong>th, Tyler – P194<br />

Daghfous, Gheylen – P93<br />

Daimon, Caitlin – 47<br />

Dale, William – P146<br />

Dalton, Pamela – P185<br />

Dana, Rachel M. – P149<br />

Davidson, Amanda J – 18<br />

de Almeida, Licurgo – P94<br />

De Araujo, Ivan – 43<br />

de Graaf, Cees – P181<br />

de Sauvage, Frederic J. – 45<br />

de Wijk, Rene A. – P181<br />

deCabo, Rafael – 47<br />

Decker, Jessica – P215<br />

Defoe, Dennis M. – P219<br />

Dekker, Teun – P187<br />

Delay, Eugene R. – P150<br />

Denman-Brice, Alexander J. – P78<br />

Denzer, Melanie Y – P133<br />

Derjean, Dominique – P93<br />

Derr, Tiffany M. – P232, P235<br />

Devaud, Jean-Marc – 53<br />

Devore, Sasha – 19, P94<br />

Dewan, Adam – 28<br />

Dewis, M.L. – P126<br />

Di Lorenzo, Patricia M – P28, P78, P85<br />

Diaz-Quesada, Marta – 30<br />

Diaz, Omar – P215<br />

131


Author Index, continued<br />

Dillon, T Samuel – 19<br />

Dlugosz, Andrezej A – P222<br />

Dobkins, Karen R. – P15<br />

Donnelly, Christopher R. – P217<br />

Doruk, Cinar – P203<br />

Dotson, Cedrick D. – P167<br />

Doty, Richard L. – P15, P21<br />

Downs, Anthony M. – P219<br />

Driver, Emily – P74<br />

Drucker-Colín, René – P18<br />

Dubuc, Réjean – P93<br />

Dudley, Jason – P74<br />

Duffy, Valerie B – P34, P231, P242<br />

Dumancic, Mira – P254<br />

Durocher, Shelley – P34<br />

Dvoryanchikov, Gennady – P151<br />

Dwyer, Patrick – P188<br />

Eckel, Lisa A. – P207<br />

Economo, Michael N – 15<br />

Einhorn, Matthew – P32<br />

El Mountassir, fouzia – P250<br />

Elson, Amanda E.T. – P167<br />

Ermakova, Irina I. – P183<br />

Ermilov, Alex – P222<br />

Escanilla, Olga – 19, P28<br />

Eslinger, Paul – P23<br />

Fadool, Debra A. – P110<br />

Fazio, Pamela – P119, P121<br />

Fedorova, Elena M. – P183<br />

Feeney, Emma L – P30, P229, P233<br />

Fei, Da – P218<br />

Feng, Guo – P37, P131<br />

Feng, Pu – P152, P200<br />

Feretic, Brian – P74<br />

Ferreira, Guillaume – P201<br />

Fiegel, Alexandra – P35<br />

Finger, Thomas E – 58, P115, P120, P123,<br />

P125, P165<br />

Finkbeiner, Susana – P234<br />

Firestein, Stuart – P91<br />

Fischer, Mary E. – P14, P184<br />

Fletcher, Max L – P88<br />

Flohr, Elena L. R. – P66<br />

Florian, Jessica – P134<br />

Fontanini, Alfredo – 33, 34, 39, P79<br />

Fooden, Andrew M. – P85<br />

Ford, Anthony P. – P165<br />

Formaker, Bradley K. – P199<br />

Forni, Paolo E. – P71<br />

Foskett, J. Kevin – 44<br />

Fradkin, Lee – 16<br />

Frank, Marion E. – P199<br />

Frasnelli, Johannes – P59<br />

Freiherr, Jessica – P57, P128<br />

Fujikawa, Toyonari – P143<br />

Fujimaru, Tomomi – P29<br />

Fujiwara, Nana – P32<br />

Fukunaga, Izumi – 38<br />

Fuss, Stefan H. – P251<br />

Gaby, Jessica M – P135<br />

Gailer, Stefan – P133<br />

Gaillard, Dany – 29<br />

Garneau, Nicole L. – P232, P235<br />

Gaudet, Rachelle – 3<br />

Gautam, Shree Hari – P95<br />

Gelperin, Alan – P202<br />

Geraedts, Maartje CP – P168<br />

Gilbertson, Timothy A. – 8, P156, P172<br />

Gire, David H. – 37<br />

Glendinning, John I. – P169<br />

Goddard, William A – 4<br />

Godfrey, Jessica – P74<br />

Golden, Glen J. – P202<br />

Goldman, Olivia – P169<br />

Gong, Qizhi – P46<br />

Gordon, Amy R – P186<br />

Gorin, Monika – P96<br />

Gotow, Naomi – P140<br />

Gottfried, Jay A. – P65, P136<br />

Grachtchouk, Marina – P222<br />

Graham, Dustin M – P210<br />

Graves, Lisa – P17<br />

Green, Amanda – P17<br />

Green, Barry G – P236, P241<br />

Green, Carter – P122<br />

Green, Erin – P174<br />

Gribble, Fiona – P110<br />

Griffith, James W. – P231<br />

Grillet, MA – P41<br />

Grosmaitre, Xavier – 48, 49<br />

Gross, Lauren A. – P225<br />

Gruss, Jason – P4<br />

132


Author Index, continued<br />

Gruver, Kim M. – P32<br />

Guarneros, Marco – P18<br />

Gumucio, Deborah L – P222<br />

Haehner, Antje – P97<br />

Haering, Claudia – P42<br />

Hajnal, Andras – P64<br />

Halabiya, Samer – P244<br />

Haley, Melissa – P79<br />

Han, Wenfei – 43<br />

Hansen, Dane R. – P172<br />

Hanson, Kyle – P106<br />

Hansson, Bill – 1<br />

Harrison, Theresa A. – P219<br />

Hassenlklöver, Thomas – P75<br />

Hatt, Hanns – P42, P44, P49<br />

Hauner, Katherina K. – P136<br />

Hayar, Abdallah M. – P204<br />

Hayes, John E – P30, P227, P228, P229, P233<br />

Hayoz, Sebastien – P72<br />

He, Jiwei – P47<br />

He, Wei – P181<br />

Hegg, Colleen C. – P72<br />

Heilman, Kenneth M – P25<br />

Henkemeyer, Mark – P215<br />

Hernandez, Sara – 43<br />

Herrmann, Christian – P44<br />

Herz, Rachel S, – P237<br />

Hettinger, Thomas P. – P199<br />

Hildebrand, John – P102<br />

Hill, David L – P210, P224<br />

Hill, Sharon R. – P187<br />

Hing, Huey – 16<br />

Hirota, Junji – P33<br />

Hirsch, Alan – P1, P2, P3, P4, P5, P9, P10, P170<br />

Hirsch, Jack – P170<br />

Hirsch, Noah – P5<br />

Hiser, Jaryd – P58, P60<br />

Hoche, Elisabeth – P59, P61<br />

Hochheimer, Andreas – P153<br />

Hoffman, Emma – 18<br />

Hoffman, Howard J. – P231, P242<br />

Holy, Timothy E – 17, P198<br />

Homma, Ryota – P89<br />

Hornung, David E – P127<br />

Hoshino, Natalia – P215<br />

Hostetler, Brian – P232<br />

Huang, Anthony – P154<br />

Huang, Guan-Hua – P14, P184<br />

Huang, Liquan – 41, P163<br />

Huang, Tao – P220<br />

Huang, Wen – P248<br />

Hudson, Robyn – P18<br />

Huerta, Ramon – P108<br />

Hummel, Cornelia – P19, P56, P63, P66, P97<br />

Hummel, Thomas – P19, P56, P63, P66, P97,<br />

P139, P145<br />

Hurst, Jane L – 18, P192<br />

Hutch, Chelsea R. – P72<br />

Hwang, Gul – P3, P4<br />

Hyder, Fahmeed – P67<br />

Iannilli, Emilia – P56, P97<br />

Ignell, Rickard – P187<br />

Inamdar, Loma B. – P234, P247<br />

Inoue, Shigeki – P43<br />

Iqbal, Tania R. – P72<br />

Irwin, Mavis A – P73<br />

Isogai, Tomoyuki – P31<br />

Ito, Kanetoshi – P143<br />

Iwatsuki, Ken – 24<br />

Jacobi, Eric – 31<br />

Jacobs, Cassandra – P164<br />

Jacobson, Aaron – P174<br />

Jaen, Cristina – P185<br />

Jahng, Jeong-Won – P211<br />

Jansen, Fabian – P44<br />

Jessica, McKinney – P11<br />

Jezzini, Ahmad – 33<br />

Jia, Cuihong – P72<br />

Jiang, Jianbo – P47<br />

Jiang, Peihua – 24, 57, P163<br />

Jiang, Yue – P53<br />

Jin, Jingwen – P137<br />

Johannes, Kornhuber – P20<br />

Johnson, Erik C – P121<br />

Johnson, Kristen – P134<br />

Jonas, Taylor – P106<br />

Jones, Joseph Wesley – P138<br />

Jyotaki, Masafumi – P162<br />

Kalbe, Benjamin – P44, P49<br />

Kang, Nana – P45<br />

Kang, Yi – P116<br />

Kanzaki, Ryohei – P43<br />

133


Author Index, continued<br />

Karunanayaka, Prasanna – P23<br />

Kass, Marley D. – 50, P98<br />

Kateri, Maria – P191<br />

Kato, Hiroyuki – P54<br />

Katz, Donald B – P80, P81, P82, P212<br />

Kaulin, Yuri – P163<br />

Kaur, Angeldeep – P191<br />

Kay, Leslie M – P203<br />

Keene, Jennifer – P212<br />

Kelber, Christina – P99<br />

Kennedy, Linda M – P244<br />

Kern, David W – 20, P133, P139, P146, P147<br />

Kharas, Natasha – P91<br />

Kilinc, Seda – P252<br />

Kim, Hyoseon – P45<br />

Kim, Jin-Young – P211<br />

Kim, MJ – P126<br />

Kim, Soo-Kyung – 4<br />

Kim, Y – P126<br />

Kimura, Meiko – P160<br />

Kinnamon, John C. – P166<br />

Kinnamon, Sue C – 58, P115, P165<br />

Klasen, Katharina – P49<br />

Klein, Barbara E.K. – P14, P184<br />

Klein, Ophir – 45<br />

Klein, Ronald – P14, P184<br />

Kobayakawa, Tatsu – P140<br />

Kochem, Matthew C – P171<br />

Koehler, Luzie K. – P19<br />

Kollndorfer, Kathrin – P59, P61<br />

Kollo, Mihaly – 38<br />

Kolterman, Brian – P100<br />

Komatsu, Yoshihiro – P221<br />

Komiyama, Takaki – 52<br />

Kondoh, Takashi – P150<br />

Koo, JaeHyung – P45<br />

Kornhuber, Johannes – P133<br />

Köster, David – P44<br />

Kotwal, Ashwin A. – P146<br />

Koulakov, Alex – P100<br />

Kowalczyk, Ksenia – P59, P61<br />

Krimm, Robin – 42, P218, P220, P223<br />

Krohn, Michael – P153<br />

Krosnowski, Kurt – P86, P101<br />

Krssak, Martin – P61<br />

Kumarhia, D – P155<br />

Kurahashi, Takashi – P54<br />

La Camera, Giancarlo – 33<br />

Lane, Robert P. – P252<br />

Lapis, Trina J. – P238, P239<br />

Larsson, Maria – P66<br />

Laska, Matthias – P18<br />

Lau, Billy Y – 35<br />

Lauer, Stephanie – P24<br />

Le Berre, Elodie – P63<br />

Le Bon, Anne Marie – P250<br />

Le Roux, Carel – 12<br />

Lee, Jacqueline – P172<br />

Lee, Jong-Ho – P211<br />

Lee, Joo-Young – P211<br />

Lee, Robert J. – 57<br />

Lee, Wooje – P46<br />

Lei, Hong – P102<br />

Leinders-Zufall, Trese – 31<br />

Lemasters, Lucas – P141<br />

Lemon, Christian H. – P36, P116<br />

Leon-Sarmiento, Fidias E. – P21<br />

Létourneau, Jean-Luc – P93<br />

Levy, Efrat – P24<br />

Lewandowski, Brian C – P163<br />

Lewis, Matthew – 19<br />

Lewis, Rosemary – P47, P76<br />

Li, Anan – P103<br />

Li, Jennifer X. – P80<br />

Li, Libo – P222<br />

Li, Wen – P58, P60<br />

Li, Yan – 24<br />

Li, Yun R. – P53<br />

Lian, Lu-Yun – P192<br />

Liggett, Stephen – 56<br />

Light, Kim E. – P204<br />

Lim, Juyun – P29, P238, P239<br />

Lin, Weihong – P50, P86, P101, P117<br />

Lin, Xiong B. – P204<br />

Lind, Nina – P114<br />

Linster, Christiane – 19, P94<br />

Lipchock, Sarah V. – P240<br />

Lisman, John – P100<br />

Liu, Fei – 29<br />

Liu, Hongxiang – P221, P222<br />

Liu, Yan – P156<br />

Livdahl, Todd P – P244<br />

134


Author Index, continued<br />

Llewellyn-Smith, Ida – P110<br />

Logan, Darren – P191<br />

Lopez-Guzman, Mirielle – P24<br />

Lopez, Fabian – P195<br />

Lopez, Roberto – P195<br />

Lossow, Kristina – P159<br />

Low, Lois – P106<br />

Lucero, Mary T – P73<br />

Lundström, Johan N – 32, P57, P62, P132,<br />

P144, P186<br />

Lundy, Robert – P182<br />

Luo, SiWei – P32<br />

Lustbader, Elias D – P246<br />

Lyall, V – P126<br />

Ma, Limei – P76<br />

Ma, Minghong – P40, P47, P76<br />

Ma, Zhongming – 44<br />

Mackay, Trudy F C – P248<br />

Maffei, Arianna – P79<br />

Magableh, Ali – P182<br />

Maier, Joost X – P81<br />

Mainland, Joel – 9<br />

Mainland, Joel D. – P148<br />

Maîtrepierre, Elodie – P161<br />

Majeed, Shahid – P187<br />

Maliphol, Amanda B – P157<br />

Manella, Laura – 19<br />

Mansfield, Corrine J. – P243<br />

Manuel, Santiago – P11<br />

Manzini, Ivan – P75<br />

Marambaud, Philippe – 44<br />

Margolis, Frank – P45<br />

Margolskee, Robert F. – 24, 53.5, 59,<br />

P163, P169<br />

Marino, Susan – P13<br />

Marrero, Yasmin U – P82<br />

Martens, Jeffrey R – P48<br />

Martin, Bronwen – 47<br />

Martin, Louis J. – P158<br />

Marton, Tobias – P191<br />

Mathew, Phoebe S. – P243<br />

Matsubasa, Tomoko – P140<br />

Matsumoto, Ichiro – P33, 44<br />

Matsunami, Hiroaki – P53<br />

Mattes, Richard D. – 7, P178<br />

Mattingly, Jameson K – P115<br />

Mattson, Sarah N. – P7<br />

Maudsley, Stuart – 47<br />

Maute, Christopher – P185<br />

Mazzucato, Luca – 33<br />

McCaughey, Stuart A. – P149<br />

McClintock, Martha K. – 20, P139, P146, P147<br />

McCrohan, C – P41<br />

McGann, John P. – 50, P98<br />

McGeary, John E – P30, P227, P233<br />

Mchugh, Robert – P64<br />

McIntosh, Elissa C. – P6<br />

McIntyre, Jeremy C – P48<br />

McKenna, Amanda K – P104<br />

McKinney, Jessica – P138<br />

McLean, Lynn – 18, P192<br />

McNamara, Patty – P232<br />

McTavish, Thomas S. – P213<br />

Medler, Kathryn F – P157<br />

Mehta, Nisarg M – P203<br />

Menefee, Abigail B. – P223<br />

Meng, Lingbin – 42<br />

Mennella, Julie A. – P234, P240, P243, P247<br />

Meredith, Michael – 6, P196<br />

Metzler-Nolte, Nils – P44<br />

Meullenet, Jean-François – P35<br />

Meyerhof, Wolfgang – 60, P159<br />

Michael, Biderman – P11<br />

Milbury, Lydia – P132<br />

Millar, Sarah E – 29<br />

Miller, Stacie S. – P144<br />

Millman, Daniel – 37<br />

Minaya, Dulce M. – P172<br />

Misaka, T – P126<br />

Mishina, Yuji – P221<br />

Mistretta, Charlotte – 23, P84, P216,<br />

P221, P222<br />

Miwa, Takaki – P70, P107<br />

Moberly, Andrew H. – 50, P98, P148<br />

Mohanty, Aprajita – P137<br />

Monk, Kevin J – P212<br />

Mony, Pauline – P35<br />

Moore, Sierra – P134, P141, P188<br />

Morgan, Charlie – P17<br />

Mosher, Zachary – P64<br />

Mosinger, Bedrich – 59<br />

Muehlberger, Andreas – P66<br />

135


Author Index, continued<br />

Müller, Christian A. – P59, P61<br />

Munger, Steven D. – 13, 22, P167, P112, P168<br />

Mura, Emi – P175<br />

Murata, Yuko – P160<br />

Murphy, Claire – P6, P7, P17, P22, P174<br />

Murphy, Melissa A – P173<br />

Murthy, Venkatesh – 34, 37<br />

Mylnikov, Sergey V. – P183<br />

Nachtigal, Danielle – P236, P241<br />

Nagaoka, Mikiya – P107<br />

Nakano, Shiori – P142<br />

Nakatani, Kei – P43<br />

Narukawa, Masataka – P159<br />

Neiers, Fabrice – P161<br />

Neuhaus, Eva – P44<br />

Ngai, John – P249<br />

Nguyen, Ha – 27<br />

Niki, Mayu – P162<br />

Ninomiya, Yuzo – P162<br />

Nitsch, Marie – P20<br />

Nixon, Ralph – P24<br />

Nomura, Shusaku – P143<br />

Noordermeer, Jasprina – 16<br />

Norbert, Thuerauf – P20<br />

Novak, Lucas R. – P58, P60<br />

Novikov, Sergey N. – P183<br />

Oboti, Livio – 31<br />

Ogg, M Cameron – P88<br />

Ogura, Tatsuya – P50, P117<br />

Ohira, Masako H. – P143<br />

Ohkuri, Tadahiro – P162<br />

Ohla, Kathrin – 32, P62, P132, P186<br />

Ohm, Daniel T. – P136<br />

Ohmoto, Makoto – P33<br />

Oke, Bukola A. – P55<br />

Oleson, Stephanie M. – P174<br />

Olsson, Mats J – P186<br />

Omelian, Jacquelyn M – P118<br />

Ortiz-Romo, Nahum – P18<br />

Osman, Allen – P21<br />

Otazu, Gonzalo H – 36<br />

Ozbek, Irene N – P11, P138<br />

Pacifico, Rodrigo – 28<br />

Pang, Peggy – P35<br />

Pankow, James S. – P14, P184<br />

Pankratz, Nathan – P14, P184<br />

Park, In Jun – P68<br />

Park, Jeeha – P34<br />

Park, Sang – P129<br />

Parker, Rockwell M. – 59<br />

Parma, Valentina – P144<br />

Pask, Gregory M – P253<br />

Paskin, Taylor R – P190<br />

Pawasarat, Ian – P21<br />

Pelchat, Marcia L. – P132<br />

Penner, Michael H. – P238<br />

Pereira, Elizabeth – P151<br />

Petersen, R – P41<br />

Peyrot des Gachons, Catherine – P175<br />

Phelan, Marie M – P192<br />

Philimonenko, Anatoly A. – P183<br />

Pierce, Dwight R. – P204<br />

Pierchala, Brian A. – P217<br />

Pinto, Alex – P14, P184<br />

Pinto, Jayant M. – 20, P146, P147<br />

Poirier, Nicolas – P161<br />

Pooley, Apryl E. – P72<br />

Prescott, John – P63, P228<br />

Preti, George – P189<br />

Primrose, Rachel J – P229<br />

Prince, Janet – P194<br />

Principe, Jose C. – P68<br />

Prokop-Prigge, Katharine A. – P189<br />

Puche, Adam C – P112<br />

Puschmann, Laura – P145<br />

Quintanilla, Carlo – P74<br />

Ramakrishnan, Vijay R – 58, P115<br />

Ramanathan, Arvind – P130<br />

Rasche, Sebastian – P44<br />

Raudenbush, Bryan – P8, P134, P141,<br />

P188, P230<br />

Rawal, Shristi – P34, P242<br />

Ray, Anandasankar – 21<br />

Rebello, Michelle R – P67, P95, P157, P213<br />

Redding, Kevin – 24, 59<br />

Reed, Danielle R. – 57, P240, P243<br />

Reimann, Frank – P110<br />

Reisert, Johannes – P202<br />

Restrepo, Diego – P103, P106, P195<br />

Reyes, Juan G – P179<br />

Reyland, Mary – 27<br />

Rhyu, MR – P126<br />

136


Author Index, continued<br />

Richards, Paige – P119<br />

Riley, Edward P. – P7<br />

Rinberg, Dima – P100<br />

Rinberg, Dmitry – 28<br />

Roberts, Kristi M. – P243<br />

Roberts, Sarah A – 18<br />

Rochlin, M William – P215<br />

Roessler, Wolfgang – P99<br />

Roeßner, Veit – P19<br />

Rogers, Ann M. – P64<br />

Rokni, Dan – 37<br />

Romagny, Sébastien – P205<br />

Romanovitch, Mallory P. – P164<br />

Roper, Stephen D. – P151, P154<br />

Rosenthal, Michelle C. – 50<br />

Ross, Brendan W. – P225<br />

Rothermel, Markus – 30<br />

Roussos, Alexander – P9<br />

Rubin, Benjamin D – P212<br />

Rucker, Joseph B. – 5<br />

Runge, Elizabeth M – P215<br />

Rupprecht, Sebastian – P64<br />

Russo, Abigail A – P82<br />

Ryan, Sarah – P23, P64<br />

Sabin, Julia – P244<br />

Sabine, Frenzel – P159<br />

Sadrian, Benjamin – P105<br />

Salcedo, Ernesto – 45, P106<br />

Sanders, Honi – P100<br />

Sandra, Hübner – P159<br />

Sanganahalli, Basavaraju G – P67<br />

Sathyanesan, Aaron – P50<br />

Saunders, CJ – P120<br />

Savigner, Agnes – P40<br />

Schaefer, Andreas – 34, 38<br />

Scheibe, Mandy – P139<br />

Schier, Lindsey A. – P176<br />

Schmaltz, Anja – 38<br />

Schmidt, Roland – P129<br />

Schmitt, Thomas – P99<br />

Schoen, Chelsea – P24<br />

Scholz, Paul – P49<br />

Schöpf, Veronika – P59, P61<br />

Schubert, Carla R. – P14, P184<br />

Schultz, Amanda – P8, P230<br />

Schumm, L. Philip – 20, P146, P147<br />

Schwab, Dieter – P20<br />

Sclafani, Anthony – P169<br />

Segil, Neil – 26<br />

Seidel, Kerstin – 45<br />

Seo, Han-Seok – P35<br />

Setlow, Barry – P209<br />

Seubert, Janina – P62<br />

Shaikh, Noorussabah – P10<br />

Sharafi, Mastaneh – P34<br />

Sharples, Ashley – P226, P245<br />

Shea, Stephen – 34, 35<br />

Shen, Kang – P194<br />

Shepherd, Gordon M – P67, P213<br />

Sherman, San<strong>for</strong>d – P3<br />

Shiga, HIdeaki – P70, P107<br />

Shoaf, Madison L – P121<br />

Shykind, Benjamin – P39<br />

Sigoillot, Maud – P161<br />

Silas, Jonathan – P21<br />

Silver, Wayne L – P119, P121<br />

Simonton, Ariel R – P196<br />

Sinakevitch, Irina – P108<br />

Sinding, Charlotte – P63, P145<br />

Small, Dana M – P177, P246<br />

Smeets, Paul AM – P38<br />

Smith, Adrian – P108<br />

Smith, Brian H – P108<br />

Smith, David W – P209<br />

Smith, James C. – P207<br />

Smith, Kimberly R – P206<br />

Smith, Steve – P165<br />

Snyder, Derek J – P12, P13<br />

Snyder, Lindsey L. – P148<br />

Sögünmez, Nuray – P251<br />

Sollars, Suzanne I – P118, P158<br />

Son, HJ – P126<br />

Song, SH – P126<br />

Soto, Michele – P170<br />

Spector, Alan C. – P176, P206<br />

Spehr, Jennifer – P193<br />

Spehr, Marc – P44, P96, P191, P193<br />

Spielman, Andrew I. – P240<br />

Stamps, Jennifer J. – P13, P25<br />

Stano, Sarah – P169<br />

Steinle, Nanette I. – P167<br />

Stephan, Aaron B – P51<br />

137


Author Index, continued<br />

Storace, Douglas A – P109<br />

Stowers, Lisa – P191<br />

Strasser, Bernhard – P61<br />

Strat<strong>for</strong>d, Jennifer M – P83, P125<br />

Strowbridge, Benjamin – 34, 40<br />

Su, Xaiorong – P122<br />

Sukumaran, Sunil K – P163<br />

Sun, Chengsan – P224<br />

Sun, Xiaoyu – P23<br />

Sun, Xue – P177<br />

Sunahara, Roger K. – 2<br />

Sung, Uhna – P109<br />

Swarup, Shilpa – P248<br />

Szajer, Jacquelyn – P7, P22<br />

Szebenyi, Steven A. – P50, P117<br />

Tabert, Matthias – P129<br />

Tabuchi, Masashi – P43<br />

Takahashi, Tatsuyuki – P168<br />

Takeuchi, Hiroko – P54<br />

Talaga, Anna K – P51<br />

Tan, Sze Yen – P178<br />

Tartaglia, Jennifer B – P122<br />

Taruno, Akiyuki – 44<br />

Tauxe, Genevieve M. – 21<br />

Tehseen, Muhammad – P254<br />

Tellez, Luis – 43<br />

Tepper, Beverly J – P122<br />

Tewalt, William – P11, P138<br />

Thiebaud, Nicolas – P110<br />

Thom, Stephen R – P124<br />

Thomas-Danguin, Thierry – P63, P201,<br />

P205, P250<br />

Thuerauf, Norbert – P133<br />

Tian, Huikai – P47, P76<br />

Tizzano, Marco – 58, P115, P120, P123<br />

Tokar, Tonya – P35<br />

Töle, Jonas – P159<br />

Tombaz, Tuce – P89<br />

Ton, Son – P215<br />

Tordoff, Michael G. – 44, P180<br />

Torregrossa, Ann-Marie – P207<br />

Touhara, Kazushige – P197<br />

Trapp, Stefan – P110<br />

Trattnig, Siegfried – P59, P61<br />

Travers, Joseph B – 46<br />

Travers, Susan P – 46<br />

Trimpe, Darcy M – P111<br />

Trowell, Stephen – P254<br />

Tsunoda, Mai – P197<br />

Tumlin, Hannah – P138<br />

Ukhanov, Kirill – P255<br />

Unger, Ewald – P59, P61<br />

Uytingco, Cedric R – P112, P168<br />

Valentine, Megan – P164<br />

Van Houten, Judith L. – P164<br />

Vandenbeuch, Aurelie – P165<br />

Vasavada, Megha – P23, P64<br />

Vasselli, Joseph R. – P169<br />

Veldhuizen, Marga G – P177, P246<br />

Ventura, Alison K. – P247<br />

Verbeurgt, Christophe – P256<br />

Verhagen, Justus V – P67, P95, P213<br />

Vesek, Jeffrey – P64<br />

Victor, Jonathan D. – P85<br />

Vigues, Stephan – P168<br />

Villar, Pablo S – P179<br />

Voigt, Anja – P159<br />

Voznesenskaya, Anna – P180<br />

Wachowiak, Matt – 15, 30, P90<br />

Wall, Crystal M. – P52<br />

Wang, Hong – P152, P200<br />

Wang, Jianli – P23, P64<br />

Wang, Min – P84<br />

Wang, Rui – 47<br />

Wang, Shih-Kai – P221<br />

Wäring, Janine – P42<br />

Warren, Craig B – P129<br />

Washiyama, Kohshin – P70, P107<br />

Weiler, Elke – P69<br />

Weiss, Michael S. – P85<br />

Weitekamp, Christopher – P23<br />

Welp, Annalyn – P119<br />

Wesson, Daniel W. – P208<br />

White, Evan J – P190<br />

White, Theresa L – P228<br />

Wieser, Matthias J. – P66<br />

Wijngaarden, Clare – P38<br />

Wilkin, Françoise – P256<br />

Willhite, David C. – P213<br />

Williams, Christopher – P151<br />

Williams, Diana L. – 14<br />

Wilson, David M. – P36<br />

138


Author Index, continued<br />

Wilson, Donald A – P24, P87, P105, P113, P201<br />

Wilson, Tamika – P185<br />

Wise, Paul M – P31, P124<br />

Wolf, Madeline – P124<br />

Wray, Susan – P71<br />

Wroblewski, Kristen E. – 20, P146, P147<br />

Wu, Keng Nei – P65<br />

Wu, Yuping – 16<br />

Xu, Jiang – 41<br />

Xu, Pei S. – 17, P198<br />

Xu, WenJin – P24<br />

Yamaguchi, Tatsuya – P33<br />

Yamamoto, Junpei – P70, P107<br />

Yang, Danica – P169<br />

Yang, Qing – P23<br />

Yang, Qing X. – P64<br />

Yang, Ruibiao – P166<br />

Yano, Junji – P164<br />

Yee, Karen – 24<br />

Yoder, Wendy M – P209<br />

Yokomukai, Yoshiko – P62<br />

Yoshida, Ryusuke – P162<br />

Yu, Congrong R – P76<br />

Yuan, Ying – P76<br />

Zabawa, Christine – 30<br />

Zayas, Vivian – P135<br />

Zhao, Haiqing – 10, P51, P52<br />

Zhao, Kai – P47<br />

Zhao, Shaohua – P76<br />

Zhou, Bin – P37, P131<br />

Zhou, Wen – P37, P131<br />

Zielinski, Barbara S – P93<br />

Zinke, Holger – P153<br />

Zoon, Harriët F.A. – P181<br />

Zopf, Yurdaguel – P20<br />

Zufall, Frank – 31<br />

zur Nieden, Anna-Nora – P57<br />

Zwiebel, Laurence J – P253<br />

139


Visual Program-at-a-Glance<br />

Posters listed <strong>for</strong> AM and PM Sessions are displayed all day long.<br />

REGISTRATION<br />

3:30 pm to 7:00 pm<br />

WEDNESDAY, APRIL 17<br />

REGISTRATION<br />

7:00 am to 1:00 pm, 6:30 pm to 7:30 pm<br />

THURSDAY, APRIL 18<br />

8:00 am<br />

8:15 am<br />

8:30 am<br />

8:45 am<br />

9:00 am<br />

9:15 am<br />

9:30 am<br />

9:45 am<br />

10:00 am<br />

10:15 am<br />

10:30 am<br />

10:45 am<br />

11:00 am<br />

11:15 am<br />

11:30 am<br />

11:45 am<br />

12:00 pm<br />

12:15 pm<br />

12:30 pm<br />

12:45 pm<br />

1:00 pm<br />

1:15 pm<br />

1:30 pm<br />

1:45 pm<br />

2:00 pm<br />

2:15 pm<br />

2:30 pm<br />

2:45 pm<br />

3:00 pm<br />

3:15 pm<br />

3:30 pm<br />

3:45 pm<br />

4:00 pm<br />

4:15 pm<br />

4:30 pm<br />

4:45 pm<br />

5:00 pm<br />

5:15 pm<br />

5:30 pm<br />

5:45 pm<br />

6:00 pm<br />

6:15 pm<br />

6:30 pm<br />

6:45 pm<br />

7:00 pm<br />

7:15 pm<br />

7:30 pm<br />

7:45 pm<br />

8:00 pm<br />

8:15 pm<br />

8:30 pm<br />

8:45 pm<br />

9:00 pm<br />

9:15 pm<br />

9:30 pm<br />

9:45 pm<br />

10:00 pm<br />

10:15 pm<br />

10:30 pm<br />

10:45 pm<br />

ACHEMS EXECUTIVE<br />

COMMITTEE MEETING<br />

12:00 pm - 3:30 pm<br />

Vista Ballroom<br />

WELCOME/AWARDS CEREMONY<br />

5:00 pm - 5:45 pm<br />

Grand Ballroom Salon D<br />

GIVAUDAN LECTURE:<br />

5:45 pm - 7:00 pm<br />

Grand Ballroom Salon D<br />

POSTER<br />

SESSION I:<br />

Multimodal<br />

Reception;<br />

Chemosensation<br />

and Disease;<br />

Olfaction Periphery<br />

8:00 am - 12:00 pm<br />

Huntington Ballroom<br />

INDUSTRY<br />

SYMPOSIUM:<br />

Taste and Smell<br />

in Translation:<br />

Applications from<br />

Basic Research<br />

1:00 pm - 4:10 pm<br />

Grand Ballroom Salon D<br />

INDUSTRY<br />

RECEPTION<br />

(Ticketed Event)<br />

4:15 pm - 6:00 pm<br />

Lighthouse Courtyard<br />

PRESIDENTIAL<br />

SYMPOSIUM:<br />

Gut Peptide<br />

Interactions<br />

Between Taste,<br />

Feeding, and Reward<br />

7:30 pm - 9:30 pm<br />

Grand Ballroom Salon D<br />

140<br />

SYMPOSIUM:<br />

THE STRUCTURAL<br />

Basis of<br />

Chemosensory<br />

Signaling<br />

9:00 am - 11:15 am<br />

Grand Ballroom Salon D<br />

REFRESHMENTS<br />

AVAILABLE<br />

10:00 am - 10:30 am<br />

Grand Ballroom Foyer<br />

REFRESHMENTS<br />

AVAILABLE<br />

2:15 pm - 2:30 pm<br />

Grand Ballroom Foyer<br />

THE BARRY DAVIS<br />

WORKSHOP:<br />

Federal Funding<br />

Opportunities <strong>for</strong> the<br />

New Investigator<br />

3:00 pm - 5:00 pm<br />

Grand Ballroom Salon C<br />

REFRESHMENTS<br />

AVAILABLE<br />

7:30 pm - 8:00 pm<br />

Grand Ballroom Foyer<br />

POSTER<br />

SESSION II:<br />

Olfaction<br />

Development;<br />

Taste CNS;<br />

Neuroimaging;<br />

Olfaction CNS<br />

7:00 pm - 11:00 pm<br />

Huntington Ballroom


Visual Program-at-a-Glance, continued<br />

Posters listed <strong>for</strong> AM and PM Sessions are displayed all day long.<br />

REGISTRATION<br />

7:30 am to 12:30 pm, 7:00 pm to 8:00 pm<br />

POSTER<br />

SESSION III:<br />

Trigeminal;<br />

Human Olfactory<br />

Psychophysics; Taste<br />

Periphery<br />

8:00 am - 12:00 pm<br />

Huntington Ballroom<br />

REFRESHMENTS<br />

AVAILABLE<br />

9:30 am - 10:00 am<br />

Grand Ballroom Foyer<br />

FRIDAY, APRIL 19<br />

PLATFORM<br />

PRESENTATIONS:<br />

Olfaction<br />

8:00 am - 9:30 am<br />

Grand Ballroom Salon D<br />

SYMPOSIUM:<br />

Stem and Progenitor<br />

Cells <strong>for</strong> Taste Buds<br />

— Development and<br />

Renewal<br />

10:00 am - 12:15 pm<br />

Grand Ballroom<br />

Salon D<br />

ACHEMS BUSINESS MEETING<br />

1:00 pm - 2:00 pm<br />

Grand Ballroom Salon D<br />

REFRESHMENTS AVAILABLE<br />

PLATFORM PRESENTATIONS:<br />

Polak Young Investigator<br />

Award Winners<br />

SYMPOSIUM:<br />

New Approaches<br />

to Physiology<br />

and Behavior in<br />

Awake Rodents<br />

8:00 pm - 10:20 pm<br />

Grand Ballroom Salon D<br />

2:30 pm - 4:05 pm<br />

Grand Ballroom Salon D<br />

CHEMOSENSORY ENTERPRISE<br />

AND MENTORSHIP ALLIANCE<br />

(ChEMA) SOCIAL<br />

REFRESHMENTS<br />

AVAILABLE<br />

7:30 pm - 8:00 pm<br />

Grand Ballroom Foyer<br />

2:15 pm - 2:45 pm<br />

Grand Ballroom Foyer<br />

5:00 pm - 7:00 pm<br />

Lighthouse Courtyard<br />

POSTER<br />

SESSION IV:<br />

Chemical Signaling<br />

and Behavior;<br />

Animal Behavior/<br />

Psychophysics;<br />

Chemosensation<br />

and Metabolism;<br />

Vomeronsasal<br />

and Chemical<br />

Communication<br />

7:00 pm - 11:00 pm<br />

Huntington Ballroom<br />

141<br />

REGISTRATION<br />

7:30 a.m. to 12:15 p.m., 5:45 p.m. to 6:15 p.m.<br />

POSTER<br />

SESSION V:<br />

Human Taste<br />

Psychophysics;<br />

Olfaction Receptors;<br />

Taste Development<br />

8:00 am - 12:00 pm<br />

Huntington Ballroom<br />

SATURDAY, APRIL 20<br />

CLINICAL LUNCHEON<br />

(Ticketed Event) — Taste Receptors<br />

in Gut and Pancreas Regulate<br />

Endocrine Function<br />

12:30 pm - 2:30 pm<br />

Vista Ballroom<br />

SYMPOSIUM:<br />

The New ‘Faces’ of Chemosensation —<br />

Utilizing Chemosensory Signaling<br />

Pathways Outside the Canonical<br />

Chemosensory Organs<br />

3:00 pm - 5:15 pm<br />

Grand Ballroom Salon D<br />

CLOSING BANQUET (Ticketed Event)<br />

7:00 pm - 9:00 pm<br />

Lighthouse Courtyard<br />

SYMPOSIUM:<br />

Experience Driven<br />

Plasticity <strong>for</strong> the<br />

Olfactory System<br />

10:00 am - 12:15 pm<br />

Grand Ballroom Salon D<br />

IFF LECTURE:<br />

Bitter Taste in Mice and Man<br />

6:15 pm - 7:00 pm<br />

Grand Ballroom Salon D<br />

PLATFORM<br />

PRESENTATIONS:<br />

Taste<br />

8:00 am - 9:45 am<br />

Grand Ballroom Salon D<br />

REFRESHMENTS<br />

AVAILABLE<br />

9:30 am - 10:00 am<br />

Grand Ballroom Foyer<br />

8:00 am<br />

8:15 am<br />

8:30 am<br />

8:45 am<br />

9:00 am<br />

9:15 am<br />

9:30 am<br />

9:45 am<br />

10:00 am<br />

10:15 am<br />

10:30 am<br />

10:45 am<br />

11:00 am<br />

11:15 am<br />

11:30 am<br />

11:45 am<br />

12:00 pm<br />

12:15 pm<br />

12:30 pm<br />

12:45 pm<br />

1:00 pm<br />

1:15 pm<br />

1:30 pm<br />

1:45 pm<br />

2:00 pm<br />

2:15 pm<br />

2:30 pm<br />

2:45 pm<br />

3:00 pm<br />

3:15 pm<br />

3:30 pm<br />

3:45 pm<br />

4:00 pm<br />

4:15 pm<br />

4:30 pm<br />

4:45 pm<br />

5:00 pm<br />

5:15 pm<br />

5:30 pm<br />

5:45 pm<br />

6:00 pm<br />

6:15 pm<br />

6:30 pm<br />

6:45 pm<br />

7:00 pm<br />

7:15 pm<br />

7:30 pm<br />

7:45 pm<br />

8:00 pm<br />

8:15 pm<br />

8:30 pm<br />

8:45 pm<br />

9:00 pm<br />

9:15 pm<br />

9:30 pm<br />

9:45 pm<br />

10:00 pm<br />

10:15 pm<br />

10:30 pm<br />

10:45 pm


A NNUAL<br />

AChemS<br />

<strong>Association</strong> <strong>for</strong> <strong>Chemoreception</strong> <strong>Sciences</strong><br />

XXXVI<br />

M EETING<br />

APRIL 9-13, 2014<br />

HYATT REGENCY COCONUT POINT | FORT MYERS, FLORIDA<br />

142


Delivering Flavour and Taste Technologies<br />

that Drive Sustainable Success<br />

www.givaudan.com

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!