116
Gustatory Information Processing By Sam Reiter B.S., Brown University, 2009 Thesis Submitted in partial fulfillment of the requirements for the Degree of Doctor of Philosophy in the Department of Neuroscience at Brown University PROVIDENCE, RHODE ISLAND MAY 2014

Gustatory Information Processing By Sam Reiter Thesis

  • Upload
    others

  • View
    6

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Gustatory Information Processing By Sam Reiter Thesis

Gustatory Information Processing

By

Sam Reiter

B.S., Brown University, 2009

Thesis

Submitted in partial fulfillment of the requirements for the

Degree of Doctor of Philosophy in the Department of Neuroscience at Brown University

PROVIDENCE, RHODE ISLAND

MAY 2014

Page 2: Gustatory Information Processing By Sam Reiter Thesis

© Copyright 2014 by Sam Reiter

Page 3: Gustatory Information Processing By Sam Reiter Thesis

iii

This dissertation by Sam Reiter is accepted in its present form by the Department

of Neuroscience as satisfying the dissertation requirement for the degree of

Doctor of Philosophy

Date_______________ _________________________

Mark Stopfer, Advisor

Recommended to the Graduate Council

Date_______________ _________________________

David Berson, Reader

Date_______________ _________________________

Leonardo Belluscio, Reader

Date_______________ _________________________

Dmitry Rinberg, Reader

Approved by the Graduate Council

Date_______________ _________________________

Peter Weber, Dean of the Graduate School

Page 4: Gustatory Information Processing By Sam Reiter Thesis

iv

CURRICULUM VITAE

May, 2014

NAME, POSITION, ACADEMIC DEPARTMENT

Sam Reiter

Graduate Student

Department of Neuroscience, Brown University,

Laboratory of Cellular and Synaptic Neurophysiology, NICHD, NIH

HOME ADDRESS

7500 Woodmont Ave, Bethesda, MD, 20814. Apt 711

EDUCATION

5/2009: BS with honors, Neuroscience, Brown University, Thesis Advisor: Mayank Mehta,

Thesis title- Hippocampal Sharp Waves: Identification and Quantification

Awards and Honors

5/2007, Undergraduate Research and Teaching Award, Brown University

5/2008, Undergraduate Research and Teaching Award, Brown University

5/2009, Richard E. Whalen Award for Undergraduate Research Excellence in Neuroscience and

Behavioral Biology, Department of Neuroscience, Brown University

5/2009-5/2014 Intramural Research Training Award (IRTA), National Institutes of Health

Teaching Experience

(2012) FAES summer course: olfactory system physiology (5 hours)

Publications

Reiter SA, Stopfer M (2013), Spike Timing and Neural Codes for Odors. In: Spike Timing:

Mechanisms and Function. (Ed. Patricia M. DiLorenzo; Jonathan D. Victor) CRC Press.

Conference Presentations

Reiter SA, Stopfer M (2010) Gustation, multimodal integration, and learning. NIH Graduate

Student Research Festival.

Reiter SA, Stopfer M (2011) Gustatory interneurons in an insect respond to multiple tastants and

with temporally complex firing patterns. Society for Neuroscience.

Reiter SA, Stopfer M (2012) Spatiotemporal coding in gustatory receptor neurons. Association

for Chemoreception Sciences.

Reiter SA, Stopfer M (2012) Gustatory information processing in an insect. Brown-GPP event at

the Marine Biological Laboratory at Woods Hole. *Won the best poster award

Reiter SA, Stopfer M (2013) Combinatorial coding of gustatory information. NIH Graduate

Student Research Festival.

Page 5: Gustatory Information Processing By Sam Reiter Thesis

v

Acknowledgments

I would not have been able to complete this project without the help of many. First there

are the people that I worked with directly. I would like to thank Kui Sun for taking care of the

animals, Chelsea Campillo for running the behavioral experiments, Tom Talbot, George Dold

and the section for instrumentation for helping design and maintain my experimental setup, Zane

Aldworth and Nitin Gupta for working on the spike sorting method, James Mcfarland for

working on neural modeling, and the entire Stopfer lab for many helpful suggestions as the work

progressed. My thesis committee of Dave Berson, Leo Belluscio and Mark Stopfer has been

consistently filled with ideas, and fun to talk with. I would especially like to thank my mentor

Mark Stopfer for giving me free reign to choose my topic of study, and for providing constant

support and encouragement throughout my time in his lab.

Equally important have been all the wonderful people in my life that have given me the

ability and moral support necessary for completing this project. I would like to thank my parents

Gail and Ira Reiter, my sister Emmy Reiter , and my friends and classmates. And perhaps most

of all, I would like to thank my wife Mari Seto for her love, and for her patience.

Page 6: Gustatory Information Processing By Sam Reiter Thesis

vi

Table of Contents

1. Background (1)

a. Neurons as letters (1)

b. Evaluating Young’s theory (3)

c. Perceptual labelled lines (5)

d. Putting Young and Müller on the same footing (7)

e. Basic Tastes (9)

f. Gustatory behavior (15)

g. Neural population activity (19)

h. The activity of single neurons (24)

i. Taste receptors can receptor cells (27)

1. Bitter (30)

2. Sweet and Umami (32)

3. Receptor addition experiments (35)

4. Sour (37)

5. Salty (37)

6. Summary (39)

j. Insect Studies (43)

2. Results (46)

a. Moth gustatory behavior (46)

b. My electrophysiological prepatation (49)

c. Profiling the responses of GRNs to tastants (57)

d. Describing the responses of GRNs over time (62)

e. Estimating the information content of the GRN population (72)

f. Second order gustatory neurons (74)

3. Discussion (83)

4. Bibliography (88)

Page 7: Gustatory Information Processing By Sam Reiter Thesis

vii

List of Illustrations

Figure 1: Tastant specific moth behavior (48)

Figure 2: A schematic of the tastant delivery system. (50)

Figure 3: Estimating the concentration of tastant that reaches the moth. (51)

Figure 4: Color sensor reading is proportional to tastant concentration. (51)

Figure 5: The maxillary nerve. (53)

Figure 6: GRN intracellular recordings. (53)

Figure 7: GRNs exhibit a diversity of tastant responses. (58)

Figure 8: A closeup of the responses of figure 7’s GRN 43. (58)

Figure 9: Summary statistics of GRN population data. (60)

Figure 10: The tastant responses of GRNs are concentration specific. (60)

Figure 11: GRNs respond to increasing concentration with increased response strength. (61)

Figure 12: The concentration dependency of GRNs is consistent across tastants. (61)

Figure 13: Trial to trial reliability of temporally rich tastant delivery. (66)

Figure 14: Comparing the rate of a GRN to that predicted under a LNM. (68)

Figure 15: The LLX improvement of a GRN model vs the response strength of the GRN. (70)

Figure 16: GRNs excibit a diversity of temporal receptive fields. (71)

Figure 17: Tastant specific information is contained in the GRN population. (73)

Figure 18: A monosynaptic EPSP seen in a spike triggered average. (75)

Figure 19: Anatomy of a SON. (76)

Figure 20: SON processes extend beyond the GRN projections. (77)

Figure 21: A narrowly tuned GRN synapses onto a broadly tuned SON. (78)

Figure 22: Summary statistics of SON population data. (79)

Figure 23: The tastant responses of SONs. (80)

Figure 24: SONs transform their input temporally. (81)

Figure 25: Tastant specific information is contained in the SON population. (81)

Figure 26: SONs classify tastants more efficiently than GRNs. (82)

Page 8: Gustatory Information Processing By Sam Reiter Thesis

1. BACKGROUND

1a. Neurons as letters

In the year 196 B.C., Egyptian priests recorded the praise of their ruler, the

Pharaoh Ptolemy 5, in stone. They carved in the two scripts of Egypt at the time, the

more formal Hieroglyphs, as well as the written lingua franca, Demotic. But they also

carved in a third language. This was because the Ptolemys, unlike earlier dynasties, were

Macedonian. Ptolemy 1st was one of Alexander the Great’s generals, helping Alexander

conquer all the lands from Greece to India. When Alexander died at age 33, there was no

one named to inherit his vast empire. The empire quickly fragmented, none of

Alexander’s generals commanding as much support as Alexander himself had. Ptolemy

1st ended taking control of Egypt, and so Ptolemy 5’ths priests carved their praise in

Greek.

In 30 BC, the first Roman emperor, Octavion, finally defeated his rivals Mark

Anthony and Cleopatra, and annexed Egypt into the Roman Empire. This ended 352

years of Ptolemaic rule. Latin became the written language, and relatively quickly

Hieroglyphs and Demotic script fell out of use and were forgotten.

Curious individuals had been trying, unsuccessfully, to decipher the meaning of

Hieroglyphs for ~1300 years when Ptolemy 5th’s stele was discovered near the city of

Rosetta by the French expedition to Egypt in 1799. It was taken by the British two years

later. By two years after that, the ancient Greek script was translated. It was quickly

recognized that the Demotic and Hieroglyphic scripts roughly corresponded to the Greek,

and that therefore this stone could be used as a code book for the forgotten languages.

1

Page 9: Gustatory Information Processing By Sam Reiter Thesis

This prospect seemed especially interesting to the British polymath Thomas

Young. Early attempts at translation were hampered by the assumption, made for as long

as people had been trying to decode hieroglyphics, that individual pictographs

represented individual words. Young realized that the word “Ptolemy” was composed of

multiple pictographs, and therefore individual pictographs represented sounds or letters,

rather than whole words. This insight, independently made by the French scholar Jean

François Champollion, was what eventually led to the translation of the Rosetta stone,

and Hieroglyphics in general.

While working on the problem of the Rosetta Stone, Young also formulated a

theory of color vision. The theory is quite concise:

“Now, as it is almost impossible to conceive each sensitive point of the retina to contain an infinite number of particles (receptors), each capable of vibrating in perfect unison (responding) with every possible undulation (wavelength), it becomes necessary to suppose the number limited, for instance (to three); and that each of the particles is capable of being put in motion less or more forcibly, by undulations differing less or more from a perfect unison; for instance, the undulations of green light will affect equally the particles in unison with yellow and blue, and produce the same effect as a light composed of those two species; ...” 1 “the different proportions, in which (the motions) may be combined, afford a variety of traits beyond all calculation.”2 Parentheses from3.

Young recognized that to account for humans’ fantastic visual acuity would

seemingly require an infinite number of receptor cell types. But this poses a problem:

there isn’t an infinite amount of space in the retina. His theory solves this problem by

hypothesizing that the retina uses a ‘population code’. Instead of many receptor cell types

each responding to a small range of wavelengths of light, Young hypothesized that there

2

Page 10: Gustatory Information Processing By Sam Reiter Thesis

could be a small number of receptor cell types each responding to a wide range of

wavelengths. By looking at the pattern of activity across the different types of receptor

cells, the brain could uniquely represent a very large number of colors without having to

build and maintain a very large number of receptor cells.

It is interesting to consider how similar Young’s theory of color vision is to

Young’s thoughts about Hieroglyphs. Translating Hieroglyphs required an understanding

that some words are built up from different combinations of multifunctional pictographs,

exactly like an alphabet. Similarly, In Young’s theory of vision different colors are

initially represented by different patterns of activity across a small number of

multifunctional receptor cell types. In both settings, combinatorics allows many things to

be represented using a few components (letters/cells). One wonders how much of

Young’s thinking about the brain was influenced by his thinking about ancient Egypt.

1b. Evaluating Young’s theory

Modern neuroscience supports Young’s theory of vision. Humans possess only a

handful of visual receptor cell types. These cells can be characterized by presenting a

range of wavelengths of light to a cell, and recording the strength of the physiological

response. This is called the cell’s action spectrum. Doing so for all visual receptor cell

types reveals overlapping action spectra each covering large ranges of the visible light

spectrum.

If these broadly responsive receptor cells were independently used to signal the

existence of certain wavelengths, the brain would be presented with an ambiguous signal.

Using information theoretic terminology, such a code would not provide enough

3

Page 11: Gustatory Information Processing By Sam Reiter Thesis

information to allow for the kinds of visual behaviors humans are capable of. We can

therefore rule out such a code, using the data processing inequality4. Evidently, these

receptor cells do not work in isolation, and precise wavelength information is extracted

by computing the pattern of activity across the different receptor cell types.

Modern neuroscience has also extended the scope of Young’s idea. Vision can be

thought of as just one place where the brain confronts a general problem: how to uniquely

represent a large number of things with a small number of cells? This is perhaps the

fundamental problem in all our sensory systems, where “things” are locations on the skin

(somatosensation), wavelengths of mechanical vibration (audition), or chemicals

(olfaction and gustation). The same problem is confronted in the context of learning and

memory, where “things” are unique memories, and in motor systems, where “things” are

precise movements. Young’s general idea provides a general answer: the solution is to

represent individual things across populations of multifunctional cells. I will hereafter

refer to this by the term ‘combinatorial code’. There is now ample evidence that

combinatorial coding is used by the brain in a wide variety of contexts5,6.

A combinatorial code can often be inferred from the fact that perceptual or

behavioral discrimination thresholds are better than theoretically possible if neurons

worked in isolation. For example, if one records from a mechanosensory cell, one can

map out set of locations on the skin that cause the cell to respond. This is termed the

cell’s spatial receptive field. If single cells were working in isolation, then stimulating

any two points that cause the cell to produce a response of equal strength would be

perceptually indistinguishable. But experiments clearly show that animals can

discriminate such stimuli. This disagreement can be solved if we consider the

4

Page 12: Gustatory Information Processing By Sam Reiter Thesis

discrimination task at the level of the neural population. The receptive fields of many

mechanosensory cells overlap such that touching a single spot of skin activates a unique

pattern of cells. Although the responses of a single cell may be equal when we touch two

nearby spots of skin, the pattern of activity across the population of cells is different, and

this makes the observed behavioral performance possible. There are many cases like this,

where to move from the activity of single neurons to explanations of behavior we require

an intermediate level of explanation, that of the neural population.

It needs to be noted that generality of Young’s idea is a weakness as well as a

strength. My interest, which I presume I share with most neuroscientists, is to provide a

mechanistic account of animal perception and behavior. To do so, it will be necessary to

describe how sensory data is initially represented by the brain, and how these

representations are transformed in order to generate behavior. While combinatorial

coding constrains these representations and transformations, it does not determine them.

For example, showing that different wavelengths of light are represented across

populations of photoreceptors does not describe how that information is transformed in

order to generate a movement towards certain wavelengths and not others. This sort of

account will require a more detailed model of how neurons respond to their inputs. We

will revisit this issue.

1c. Perceptual labelled lines

Young’s theory of color vision seems all the more remarkable considering the

state of neuroscience at the time. Philosophers had been theorizing about the problem of

how we become aware of the external world ever since Aristotle. Until the 19th century,

5

Page 13: Gustatory Information Processing By Sam Reiter Thesis

it was thought that objects in the world emitted some sort of ghostly essence, and humans

possessed a 'sensorium' that was able to perceive these essences and thus become aware

of the sensory world.

19th century thinkers broke from this tradition, and began to modularize the

senses. The most famous move in this direction was made by Johannes Müller, in his law

of specific nerve energies. Müller was the first to explicitly divide the single sensorium

into five sensory systems, the now familiar vision, audition, somatosensation, gustation,

and olfaction. Using psychophysical evidence like the fact that pushing on your eye

produces a visual sensation, Müller's argued that each of the 5 senses has a sensory

surface that is most sensitive to certain types of stimulation, and that activation of one of

these sensory surfaces is somehow linked to perception of that form of energy.

“The same cause, such as electricity, can simultaneously affect all sensory organs, since they are all sensitive to it; and yet, every sensory nerve reacts to it differently; one nerve perceives it as light, another hears its sound, another one smells it; another tastes the electricity, and another one feels it as pain and shock. One nerve perceives a luminous picture through mechanical irritation, another one hears it as buzzing, another one senses it as pain. . . He who feels compelled to consider the consequences of these facts cannot but realize that the specific sensibility of nerves for certain impressions is not enough, since all nerves are sensitive to the same cause but react to the same cause in different ways. . . (S)ensation is not the conduction of a quality or state of external bodies to consciousness, but the conduction of a quality or state of our nerves to consciousness, excited by an external cause.”7

Boring sums up Müller's law by saying that he replaced “the animal spirit

explanation for 5 animal spirits, one specific to each of the 5 senses.”8 This is perhaps a

bit unfair. Müller's theory can be criticized for its vagueness, as there is no explanation

for how the different sensory surfaces become activated, or how this activation is linked

to perceptions and behaviors. But explicitly assigning different roles to our different

6

Page 14: Gustatory Information Processing By Sam Reiter Thesis

sensory organs has been hugely influential. Indeed, we still largely divide our senses into

5 categories, or ‘systems’.

Müller's theory motivated a generation of psychophysicists. The strategy taken by

the majority of workers in this area was to attempt to subdivide the senses beyond

Müller's initial partition. Using a variety of techniques, perceptual experiments were

performed on humans to try to discern how many unique sensory experiences are

possible within each of the 5 major senses. If the major senses could be subdivided in this

way, it could be argued that there was some link between the differential activation of a

sensory surface and a unique perceptual. This term created to describe this link was

‘labeled lines’. Different experimenters developed (widely) different numbers of these

perceptual labeled lines9.

1d. Putting Young and Müller on the same footing

Sensory neuroscience has been strongly influenced both by Müller’s labeled lines

and by Young’s combinatorial coding ideas. As discussed above, Young’s population

coding has been applied meaningfully to populations of neurons in all 5 of Müller’s

senses, as well as to neurons related to movement, memory, and more general settings of

information processing in neural networks.

The concept of a labeled line had to be significantly changed for it to work in

neuroscience. Unlike Young, Müller and his school were theorizing about the number of

unique perceptions, and were not directly concerned with how neurons actually converted

activation of a sensory surface into a perception. Because of this, any theory of neural

coding, including Young’s combinatorial coding, is consistent with Müller’s perceptual

7

Page 15: Gustatory Information Processing By Sam Reiter Thesis

labeled lines. However, the idea of labeled lines became associated with a different

theory of neural coding than Young’s, one more in the spirit of Müller’s work. After

observing that the activity of a single type of neuron responds to certain behaviorally

relevant stimuli, it is often posited that the perception of and behaviors related to those

stimuli can be directly linked to this single type of neuron. By this reasoning, the

perceptual labeled lines of Müller and his followers can be mapped onto single neurons

dedicated to single stimuli10.

This formulation of labeled lines, which I will call single neuron coding, as it is

really a hypothesis about how stimuli are encoded within groups of neurons, is directly

opposed to Young’s idea of population coding. In a population code a single neuron does

not represent a single stimulus; it participates in the representation of many stimuli which

are uniquely represented at the population level. In single neuron coding, a single neuron

represents a single stimulus; a range of stimuli are represented by a range of cells

sensitive to non-overlapping stimuli.

Also as we have discussed, the utility of a combinatorial code is the ability to

represent many stimuli with few neurons. In a single neuron code, representing N things

requires N types of neurons. This fact rules out single neural codes in any situation where

a large number of stimuli need to be behaviorally discriminated using a relatively small

number of neuron types. This is the regime in which most human sensory systems

operate, for example, where a spectrum of colors are represented by 3 photoreceptor

types, and over one trillion odors can be perceived11 through the activity of less than one

thousand types of olfactory receptors12.

8

Page 16: Gustatory Information Processing By Sam Reiter Thesis

Single neuron codes do appear to apply in more specialized roles. For example, in

insects certain ecologically relevant odors are perceived by receptor neurons sensitive to

only a single chemical13,14. Perhaps having dedicated neurons in these situations allows

for more sensitivity than representation through a combinatorial code would allow.

However, single neuron codes have been argued to apply in many contexts where

behavior necessitates a combinatorial code, or where the evidence is equivocal. Mcllwain

speculated that neuroscientists have been drawn towards explaining sensory systems in

terms of single neuron codes because of their simplicity, and the fact that until recently

physiologists were limited to record one neuron at a time.

“As you sit in a darkened laboratory with your attention riveted to the sounds of the audiomonitor and probe a neuron's receptive field with a tiny visual stimulus, it is easy to forget that the cell you are listening to is but one of many that are responding to the stimulus. And from there it is but a small step to the assumption that the cell is a 'labeled line' for that aspect of the stimulus which produces the most vigorous response.”5

Reflective accounts like Mcllwain’s, coupled with the advent of technologies that

allow for the activity of many neurons in a population to be recorded at the same time,

have convinced most neuroscientists that all but one of Müller’s five senses use a

combinatorial code to represent spectra of sensory information.

1e. Basic Tastes

The only exception is the sense of taste, or gustation. Here it is widely accepted

that in animals as diverse as flies and humans the gustatory system is built to recognize a

small number of ‘basic tastes’ (BTs)10. What are BTs, and how well do they describe

gustatory information processing?

9

Page 17: Gustatory Information Processing By Sam Reiter Thesis

Humans have been classifying the number of taste perceptions for quite a long

time. By the second century BCE, the Chinese recognized five tastes, sweet, sour, salty,

bitter, and acrid15. In 350 BCE. Aristotle proposed 7 tastes, sweet, succulent, pungent,

harsh, sour, saline, and bitter16. Words for three to five taste categories can be found in at

least “Afrikaans, Arabic, Albanian, Chinese, Danish, Dutch, English, French, German,

Greek, Hebrew, Hindi, Hungarian, Indonesian, Irish, Italian, Japanese, Latin, Maori,

Nepali, Oromo, Papua, Persian, Polish, Portuguese, Russian, Samoan, Sanskrit, Scotch,

Serbo-Croatian, Slovak, Spanish, Swahili, Swedish, Tagalog, Urdu, Vietnamese, and

Yiddish.”3

The presence of four taste words in German (süß –sweet, sauer-sour, salzig –salty,

bitteren-bitter) led some psychophysicists following Müller to argue that there are four

unique taste perceptions. For example, von Skramlik reasoned

“The number of sensory qualities that belong within the taste domain is naturally infinitely large; nevertheless it cannot be denied that the same sensations, even occurring from various different stimuli, reoccur quite often. From the outset, it appears that a large number of descriptions for taste sensations are not necessary. A closer inspection of the relationships shows nevertheless that the number of expressions for taste sensations is extremely insufficient. Very few adjectives are available for description, the primary being sweet, salt, sour, bitter. These adjectives are so characteristic that with practice in using these descriptions a taste quality can be represented within reasonable limits.”17

Is this reasoning valid? It is perhaps tempting to think that the presence of a small

number of taste category words, which further are largely consistent across languages,

reflects the inherent capacities of the human gustatory system. However, ignoring the

impact of culture on forming these words, as von Skramlik did, is a mistake. Studies of

human color perception sheds light on this issue18. Like tastes, most human cultures have

10

Page 18: Gustatory Information Processing By Sam Reiter Thesis

developed a small number of ‘basic color’ words (black, white, green, blue, yellow, red,

etc.) and these words are largely consistent across cultures. It has been argued, most

influentially by Berlin and Kay, that this reflects the underlying consistency of the human

visual system19. However, more recent work supports the idea that the presence and

consistency of these words instead reflect a shared cultural representation of color20.

Obvious support of this position comes from the fact that humans are able to perceive and

discriminate between many more colors than the basics. Further, genetic variation gives

rise to humans with a wide range of color discrimination ability, and yet does not result in

these peoples’ inability to use the basic color words correctly, or the desire to add new

basic color words. Surely biology constrained the development of basic color words, but

these words, like all words, were developed to describe culturally meaningful

phenomena, not physiology.

Analogously, the choice of taste words was almost certainly made based on the

kinds of chemicals that were used by most early human groups. Many cultures hit upon

souring foods by fermentation, sweet fruits grow around the world, etc. Because taste

words are the product of a complicated interplay between biology and culture they cannot

be used to reason about the information processing capacities of the gustatory system.

von Skramlik’s reasoning was mistaken.

More careful psychophysics work suggests that gustatory perception is not

accurately described in terms of 4 BTs. Addressing this point requires a clear definition

of what exactly having 4 BTs would mean for human perception. Schiffman and Erickson

propose a working definition: If humans perceive 4BTs then 4 tastes, and only 4 tastes,

are perceptually distinct and separable21. This can be contrasted with the possibility that

11

Page 19: Gustatory Information Processing By Sam Reiter Thesis

there is a continuum of taste perceptions intermediate to the 4BTs, the possibility that

there are additional tastes that are not able to be described in terms of the 4BTs, and the

possibility that there is both a continuum of tastants and additional tastes.

When a single tone is presented to human subjects, almost 100% of people

describe the perception as “singular”. When two to four tones are presented together, it is

recognized as a mixture of pure tones, and perceived as more “complex” than single

tones. This has led audition to be described as an “analytic” sense; the perception of a

combination of tones can be analyzed to uncover the component, “basic” tones22. When

individual tastants are presented to subjects, they are perceived as ‘singular’. When two

to four tastants are presented in a mixture however, most mixtures are also perceived as

singular, and of no more complexity than single tastants. Gustation can be described as a

“synthetic” sense, where mixtures of tastants produce unique perceptions that cannot be

broken down into a description of their components22. This predicts that if one uses the

4BT terminology to describe the perception of a tastant, that description will be relative

to the presence of other tastants. Indeed, NaCl (salty) is often confused with HCl (sour)

when presented in a sucrose (sweet) solution21. These findings support the idea that there

is a continuum of gustatory perceptions. Additional support comes from the fact that

many chemicals are described in terms of multiple basic tastes. Urea and MgSO4 taste

sour-bitter23. Aspartame and sodium saccharin taste bitter-sweet24.

One might concede that the taste of individual chemicals as well as mixtures form

a continuum, but assert that the 4BTs are useful in defining that continuum. This position

is well put by Spector and Kopka.

12

Page 20: Gustatory Information Processing By Sam Reiter Thesis

“…if discriminability between two bitter-tasting compounds is displayed, it could be because one or both of the compounds stimulates receptor processes that lead to other qualitative taste sensations (e.g., sweetness). For example, humans report that saccharin has both bitter and sweet taste qualities.” 25

This position treats the 4BTs as basis vectors defining a 4 dimensional perceptual

space. Chemicals and mixtures fall at various points within this space, and can be

described in terms of the 4BT basis. Specifically, we could describe the proposed space

as (sweet, sour, salty, bitter). Sucrose, the classic sweet tastant, would fall at (1,0,0,0).

Quinine, the classic bitter tastant, would fall at (0,0,0,1). Saccharin, described as bitter-

sweet, might fall at (1, 0, 0, 1).

Some reflection reveals that this actually undermines the position of the 4BTs. If

we select 4 tastants associated with multiple BTs, then as long as at least one of our

chosen tastants has some component of each BT, simple linear algebra lets us rotate the

basis of the perceptual space. For example, we could define the same space in terms of

(bitter-sweet, sour-sweet, salty-sour, and bitter-sour). In this rotated space the bitter-sweet

taste of saccharin becomes a “basic taste”, and the sweet taste of sucrose would be

considered in terms of multiple BTs. There are an infinite number of ways that the space

could be rotated, and therefore the choice of BTs becomes arbitrary. This of course

matters for gustatory psychophysics. But it should also matter to neuroscientists. If we

accept that humans can perceive a continuum of tastants, there is no more reason to look

for the neural basis of the perception of sweet as there is for the perception of sweet-sour,

or bitter-sour, or sweet-salty-bitter.

Psychophysical studies also call into question the dimensionality of tastant space.

This line of work was started by von Skramlik himself, who tried and failed to recreate

13

Page 21: Gustatory Information Processing By Sam Reiter Thesis

the perception of a tastant in terms of a mixture of sucrose (sweet), NaCl (salty), HCl acid

(sour), and quinine (bitter)17. Schiffman et al. asked subjects to describe tastants in terms

of a wide variety of taste words, including the 4BTs. They found that individual salts

were reliably described as distinct, but could often not be put in terms of the 4BTs. When

they could, the salts could be described in terms of BTs other than salty. For example, a

low concentration of NaCl is commonly perceived as sweet26,27. When subjects were

asked to rate tastants in terms of ‘sweet’, ‘salty’, ‘sour’, ‘bitter’, and ‘amount left over’,

the amount left over was often very high21.

When two chemicals typically associated with different basic tastes are mixed

together, the perception of the mixture depends on the specific chemicals used, even

when matched for perceived intensity. For example, mixing NaCl (salty) with urea (sour,

bitter) produces a taste described as less bitter than urea alone. Mixing NaCl (salty) with

MgSO4 (sour, bitter) produces a taste described as equally bitter, but less salty than NaCl

alone24,28. This suggests that the perception of a tastant cannot be described in one

dimensional terms (like bitter intensity). In order to predict how a mixture of two tastants

will be perceived it is insufficient to know how bitter urea tastes. One also needs to know

where urea falls on at least one additional dimension, the interaction with NaCl.

But perhaps the best example that taste perception has more than four dimensions

is the story of umami. ‘Umami’ is a Japanese neologism coined by Kikunai Ikeda in

1901. It combines the word umai (delicious) with mi (taste). Two staple Japanese foods,

kombu (a type of seaweed) and dashi (dried bonito flakes) were found by Ikeda to

contain high concentrations of glutamate. He argued that the perception of this taste was

distinct from the 4BTs, and therefore constituted a fifth BT. This has been largely

14

Page 22: Gustatory Information Processing By Sam Reiter Thesis

accepted by modern gustatory researchers. Humans do taste glutamate distinctly29, and

therefore umami should be recognized as a BT under our working definition. What isn’t

clear is why all the other tastants that humans can perceive as distinct aren’t given the

status of BTs.

In sum, a modern understanding of anthropology undermines von Skramlik’s

logic for initially proposing the 4 BT model, and modern psychophysics suggests that the

perception of taste forms a high dimensional continuum. Describing tastes in terms of 5

BTs does not do justice to our own capabilities. We can taste things intermediate to and

beyond the basics. This point is of course obvious to the wine or beer connoisseur, or the

chef, all of whom use much larger taste vocabularies. The scientific interpretation of this

ability has been that in everyday life the complicated perception of flavor is a

combination of a simple gustatory system and a complex olfactory system30. No doubt

the sense of smell is both complex and an important component of flavor perception.

However modern studies of gustatory perception control for this factor, and still find a

description in terms of BTs inadequate. Apparently, the sense of flavor is a multisensory

experience using high dimensional representations of both odors and tastants.

1f. Gustatory behavior

Pfaffman, while a student of Lord Adrian, was the first to study the neurons of the

gustatory system. In a fateful decision, he decided to accept von Skramlik’s assertion that

the gustatory system of humans is responsible for detecting only 4 basic tastes.

“Classical experiments on the human have led to the view that there are four basic taste qualities, saline, sour, bitter and sweet (v. Skramlik,'26). These are said to depend on the

15

Page 23: Gustatory Information Processing By Sam Reiter Thesis

existence of four different types of taste end organ, each type specifically sensitive to salt, acid, quinine and sugar respectively. The present experiments have been concerned with the search for some analogous set of chemically sensitive endings in the cat's tongue.”31

Here Pfaffman extended von Skramlik’s 4BT (this is before the idea of umami

caught on) model of perception from human to other animals. This was done without

evidence or argument. But to be fair, he proposed the theory before much evidence

existed. Contemporary studies of animal behavior reveal that there is little evidence to

support such an extension of the BTs.

The BT model has difficulty explaining the wide variety of species and tastant

specific gustatory behaviors. While it is widely believed that all animals are attracted to

things humans consider sweet, and are averse to tastants humans consider bitter, this is

not actually the case. There is no simple way to judge whether an animal will be attracted

or averse to a tastant. For example, humans, as well as rats, spiny mice, and gerbils are

attracted to the sugar maltose. Dogs, armadillos, cows, and squirrel monkeys are

indifferent or averse to maltose32. Cats are not attracted to any sugars32, and do not

discriminate it from water33. Moths will visit flowers possessing nectar high in sucrose,

while bats and flies will visit nearby flowers with glucose and fructose rich nectar34. Rats,

hamsters, gerbils, and spiny mice prefer the taste of polysaccharides to sugars. Humans

describe the taste of polysaccharides as bland and slightly unpleasant32.

Humans describe the taste of many alkaloids and glycosides as “bitter” and

unpleasant. But many animals actually prefer low concentrations of certain deterrent

chemicals over water. Goats (Capra hircus) prefer quinine dihydrochloride. Pygmy goats

(C hircus), Virginia opossum (Didelphus virginiana) and the mouse Peromyscus aztecus

prefer quinine hydrochloride. Squirrel monkeys prefer quinine sulphate. The domestic

16

Page 24: Gustatory Information Processing By Sam Reiter Thesis

mouse prefers sucrose octaacetate35. These paradoxical preferences have been given the

wonderful name of the Schweppes effect, after the brand of tonic water that humans

enjoy because of its low concentration of QHCl35.

Why is there such diversity in how different species react to tastants? Glendinning

has advanced an interesting hypothesis concerning bitter tastes36. He reasoned that every

species is presented with a dilemma: Given that some, but not all, bitter tastes are toxic,

choosing to ignore the bitter taste risks poisoning. But especially when food is scarce,

avoiding all bitter tastes risks starvation. The percent of food containing toxic chemicals

depends on what a species consumes: while nearly 10% of terrestrial plants contain toxic

chemicals37, poisonous animals are relatively rare and often signal their toxicity36. This

leads to different kinds of animals choosing different strategies.

“It is hypothesized that mammals in different trophic groups have evolved different strategies for coping with the unpredictable bitterness/toxicity relationship. Carnivores, which rarely encounter bitter and potentially toxic foods, have a low bitter threshold and tolerance to ingested poisons; herbivores (particularly browsers), which commonly encounter bitter and potentially toxic foods, have a high bitter threshold and tolerance to ingested poisons; and omnivores, which encounter bitter and potentially toxic foods somewhat less frequently than herbivores, have an intermediate bitter threshold and tolerance to ingested poisons.” 36.

This theory does an admirable job of relating the study of gustatory behavior to

evolutionary theory. It further provides a useful framework for thinking about other

differences in taste preferences across trophic groups. But while Glendinning’s theory

suggests why different trophic groups have different relationships to the same tastants, it

does not explain the diversity of species-tastant relationships seen within a tropic group.

Rabbits and hamsters are both herbivores, but up to 100 mM NaCl is attractive to rabbits,

but aversive to hamsters33. A second look at the taste preferences of different species

17

Page 25: Gustatory Information Processing By Sam Reiter Thesis

described above shows that this variability cannot be explained only through differences

across trophic levels.

Instead of interpreting animal preferences in terms of the BTs, this variability can

be explained if one considers tastants individually. This extension of Glendinning’s

theory hypothesizes an evolutionary arms race between predators and prey. Prey plants

and animals are under evolutionary pressure to develop toxic chemical defenses, or to

develop nontoxic tastants that taste enough like toxic tastants to be avoided by predators

(a kind of gustatory Batesian mimicry). Predators are under pressure to develop the

ability to discriminate the truly toxic tastants from the mimics. Just like in the original

theory, the strength of this pressure will depend on a species’ trophic level. But the

diversity of species-tastant relationships comes not just from this fact, but from the fact

that these evolutionary arms races are local. Prey plants and animals are not evolving

tastant defenses to defend against all predators, just the ones in the local ecosystem.

Predators are not evolving the ability to discriminate all deterrent tastants, just the ones

found at certain levels of the local ecosystem.

This theory can be usefully applied to attractive tastants as well. Indeed, our

current understanding of nectar-pollinator relationships is chemical specific. Different

plants use different chemicals as nectar rewards to attract different local pollinator

species, which in turn will only visit certain species of plants34. Thus, the diversity of

ways in which species react to tastants can be accounted for simply through evolutionary

adaptive radiation.

18

Page 26: Gustatory Information Processing By Sam Reiter Thesis

1g. Neural population activity

Let us return to Pfaffman. Understandably ignorant of later work on animal

behavior, Pfaffman developed a preparation in the cat (later, hamster) where one of the

afferent nerves connecting the tongue to the brain could be recorded from

electrophysiologically. He delivered what he considered exemplar chemicals from the

4BT categories: sucrose for sweet, quinine for bitter, HCl for sour, and NaCl for salty.

Interestingly, he found that single neurons increased their rate of action potential (spike)

generation in response to more than one of the four chemicals. From this data, he

concluded that the BTs are each represented using overlapping sets of neurons responsive

to more than one basic taste. This has become known as the Across Fiber Pattern theory

in gustation.

“The extension of Muller's law of specific energies to cover different qualities of sensation within the different modalities equates phenomenological classes with anatomical or physiological entities. As experience and evidence in the field of physiological psychology generally accumulate, it is apparent that such one-for-one correlations of physiological with psychological terms cannot be supported. Thus the patterns of sensitivity found in the electrophysiological studies do not conform to the four basic taste categories of salt, sour, bitter, and sweet.”38

While superficially similar, this theory is not the same as Young’s combinatorial

coding. In Pfaffmann’s theory, the goal of the gustatory system is to encode BT

categories. In Young’s theory, the goal is to encode a large number, or even a spectrum

of sensory stimuli. Both theories argue that their respective goals are achieved using

neurons sensitive to multiple stimuli. But while Young reasoned that broadly tuned

neurons are necessary in order to encode a large number of stimuli using a small number

19

Page 27: Gustatory Information Processing By Sam Reiter Thesis

of neurons, it is unclear why the brain would need to use broadly tuned neurons if there

were only 4-5 stimuli to encode, as in Pfaffmann’s theory.

The AFP theory proved to be quite influential. In the years following Pfaffmann’s

first recordings, a collection of gustatory areas were discovered in the brain, and most of

these areas have been studied physiologically. The CT nerve that Pfaffman recorded from

is but one of three afferent nerves involved in vertebrate gustation, the others being the

greater superficial petrosal nerve and the glossopharyngeal nerve. 10 Different nerves

innervate different but overlapping regions of the tongue and soft palate, receiving input

from taste receptor cells. All the nerves project to the nucleus of the solitary tract (NST)

in the brainstem39. This nucleus also receives somatosensory input through the lingual

branch of the trigeminal nerve40. It connects with oral motor nuclei of the brainstem, as

well as with the ipsilateral pontine parabrachial nucleus (PBN). The PBN makes

reciprocal connections with the lateral hypothalamus, central nucleus of the amygdala,

bed nucleus of the stria terminalis, and insular (gustatory) cortex41. It also projects

bilaterally to the ventral posterior medial (VPM) nucleus of the thalamus, which in turn

makes reciprocal connections with gustatory cortex and the amygdala42.

Early physiological work closely followed Pfaffmann’s method: One of the above

mentioned nerves or nuclei was targeted with extracellular electrodes in an anesthetized

animal. A small number of tastants, chosen as exemplars of the BTs, were washed over

the animals tongue for 5-10 seconds, and the number of spikes during that period was

counted. Neurons responsive to multiple stimuli were found at all levels of the gustatory

system in a variety of animals, including primates43–52.

20

Page 28: Gustatory Information Processing By Sam Reiter Thesis

These results established some of the main pieces of the gustatory system, and

were consistent with the AFP theory. It is difficult to conclude much about how tastant

information is processed from these studies, because of the small number of stimuli used

and the crude way of describing neural activity. Indeed, by using only single examples of

the BTs, these studies make it impossible to distinguish whether the gustatory system

uniquely represents tastants through combinatorial coding, or BT categories through an

AFP code.

Later work has employed more sophisticated methods for analyzing neural

responses. The early practice of counting the number of spikes in the 5-10 seconds

following tastant stimulation can be seen as a strongly biased hypothesis about how

gustatory neurons represent information. If downstream neural circuits are sensitive to

this statistic of spiking, then all is well. But if downstream neurons are sensitive to other

statistics, then it is possible that information used by the brain is missed by the choice of

analysis.

When the tastant responses of neurons from many levels of the gustatory system

were examined at a finer temporal scale, it was found that single neurons responded with

different temporal patterns of activity to different tastants53. This has led some to

hypothesize that the BTs are represented by broadly responsive neurons producing

category specific temporal patterns of spiking activity54–60.

Additional support for considering the more detailed temporal dynamics of

gustatory neural activity comes from animal behavior. Animals have been shown to

detect61 and discriminate62 tastants in under a second. When compared to NST neural

21

Page 29: Gustatory Information Processing By Sam Reiter Thesis

responses from a population of neurons, discrimination time correlates with the

difference in spike rate at this fast timescale62.

Other later work used increasingly large sets of tastants to characterize gustatory

neurons. These studies showed that not only do neurons at many levels of the gustatory

system respond to multiple BTs, but neurons are differentially sensitive to tastants

associated with a single BT. For example, Dahl et al. 63 presented several tastants

considered bitter to rats while recording from two of the nerves connecting the tongue to

the brain. They found one neuron that responds to denatonium(d) and strychnine(s), but

not caffeine(c) or yohimbine(y). Another neuron responds to d,c, and y, but not s.

Another responds to c, but not d,c, or s. Another responds to d,c,y and s, etc63.

Researchers confronted these tastant specific neural responses in one of two ways.

Some sought to reduce the observed diversity of neural types. This was often done using

a statistical technique called hierarchical clustering. In this method, every neuron is

represented as a vector with length equal to the number of tastants delivered, and values

equal to the number of spikes that that the neuron produces in response to each tastant

(typically in the 5-10 s following tastant delivery). Each neuron-vector is first considered

as a unique ‘cluster’. The two closest clusters, under a Euclidean distance metric, are then

clustered together, followed by the next nearest, and then the next, etc. This process ends

when all neuron-vectors are clustered together.

By setting a threshold on the maximum distance allowed to group two clusters,

neurons can always be grouped into a small number of categories. The response profiles

of individual neurons within a category are then averaged together, and the distinguishing

features of the cluster-average responses discussed. As the averaged responses of

22

Page 30: Gustatory Information Processing By Sam Reiter Thesis

different clusters respond to stimuli associated with more than one basic taste, this

analysis is usually interpreted as supporting an AFP code for BT category64–74.

The practice of identifying a small number of clusters in hierarchical clustering

and ascribing significance to them represents a misuse of the technique75,76. Without prior

information motivating the selection of a certain distance threshold and thus number of

clusters, the choice is an arbitrary one. Clustering and then averaging the responses of

different neurons in a cluster serves to reduce the heterogeneity of neural responses,

creating only the illusion of simplicity.

Other researchers embraced the heterogeneity of neural responses, realizing that it

could allow for tastant specific, rather than BT category specific information to be

represented by the gustatory system63,77–79. Woolston and Erickson put things clearly:

“There seems to be neither compelling empirical nor theoretical evidence for ignoring the diversity seen in the response profiles of taste neurons in favor of classification schemes. The data do not force the conclusion of response pattern types, and the logic for such types seems equivocal.”80

Recently, several studies have combined a large stimulus set with more detailed

data analysis. Tastant specific temporal responses have been observed in several

populations of neurons81,82. For example, Wilson et al. recorded neural activity from the

NST of mice being presented with multiple tastants considered bitter by humans.

Different neurons were found to not only respond to different but overlapping sets of

tastants, but also to respond with tastant specific temporal patterns of activity. Studying

the evolution of the population activity over time, the representation of different tastants

could be pictured as population vectors tracing out tastant specific trajectories through

neural state space. This suggests the possibility that postsynaptic neurons use the

23

Page 31: Gustatory Information Processing By Sam Reiter Thesis

representation of different tastants across neurons and time to discriminate tastant

identity, a specific type of combinatorial coding82.

1h. The activity of single neurons

30 years after producing the AFP coding theory, Pfaffman, working with his

student Frank, reinterpreted their data. While hamster CT neurons increased their spike

rate to more than one of the 4BT exemplar chemicals, they all respond most strongly to a

single chemical. In 1973 Frank put forward the argument that these ‘best’ responses are

what the neurons actually represent47. The observed responses to ‘non-best’ stimuli were

described as merely noise, presumably removed by subsequent populations of neurons.

She concluded that the gustatory system uses a single neuron code to represent the 4BTs

(see page 7 for a discussion of single neuron codes. This theory is usually referred,

confusingly, as gustatory labelled lines).

“In the more recent results on gustation in our laboratory and those of others, although individual receptors appear to have a wide sensitivity, i.e. a chemical spectrum, they are classifiable into receptor types or clusters. These can be identified and designated by the chemical which causes the largest response, i.e. their best stimulus. In the early days of taste electrophysiology, these apparently more complex relationships caught our attention almost to the point of obscuring the regularities present.”83

Researchers following Frank’s early work conducted many of the same type of

experiments described above: Tastants representative of different BTs were presented to

an animal while recording from a gustatory nucleus or nerve. The spikes occurring in the

5-10s following tastant presentation were counted, and the taste that elicited the most

spikes became the recorded neuron’s “best” taste. This “best” taste is interpreted as what

24

Page 32: Gustatory Information Processing By Sam Reiter Thesis

the neuron really is representing; all “non-best” responses were considered to be

“noise”74,84–87.

These studies suffer from the same limitations as the early AFP studies, namely a

stimulus set that prevents the discrimination of BT category vs. tastant specific neural

responses, and a crude encoding model. In addition, they have a pretty large logical

problem built in to the experimental interpretation: the fact that regardless of the set of

stimuli and the metric used for determining response strength, there will always be one

stimulus that produces the “best” response. If the presence of a “best” stimulus is a

sufficient condition for concluding a single neuron code, as this work posits, than we

must conclude that every sensory system works through a single neuron code. This

conclusion is ruled out by information theory. As I discussed when introducing

combinatorial codes, it is possible to rule out single neuron codes in the early visual or

somatosensory systems based on the high information content of animal behavior, and

the low information content of single neuron codes. Therefore, the presence of a “best”

stimulus cannot be used as sufficient evidence for concluding that a system employs a

single neuron code. While this is the case regardless of the distribution of response

strengths, the fact that responses to “non-best” stimuli are often nearly as strong as “best”

stimulus responses47 casts this theory even further into doubt.

Although partitioning responses into “best” and “non-best” reveals little about

gustatory information processing, the idea that some features of a neuron’s activity are

used by postsynaptic neurons while others are filtered out is an interesting one. Neurons

are certainly capable of producing spiking responses that reflect a nonlinear combination

of their inputs. Because of this, it is very possible that a neuron could be sensitive only to

25

Page 33: Gustatory Information Processing By Sam Reiter Thesis

the very strongest responses of their presynaptic inputs. However, if this actually

occurred in the gustatory system, one would expect neurons at higher levels of the system

to respond to fewer tastants than their presynaptic neurons. Postsynaptic neurons would

filter out their weaker inputs, and respond only to those tastants that activated presynaptic

neurons the strongest. Synaptically connected paired recordings, which could test this

idea directly, have not been performed. But the fact that neurons in tongue, the cortex,

and the areas in between respond broadly to many tastants suggests that such a filtering

operation is not performed by the gustatory system.

To summarize, the idea of BTs has had a profound influence on physiological

studies of the gustatory system. The two major theories, AFP coding and single neuron

coding, are both hypotheses on how the BTs are represented by the system, and do not

question the validity of the BTs themselves. The diversity of neural responses seen at

many levels of the gustatory system does not lend itself well to an interpretation in terms

of the BTs. The observed diversity is expected if the system employs a combinatorial

code for individual tastants.

The influence of BTs on this area have caused many to focus on whether neurons

respond to certain types of tastants and not to others, rather than focusing on how any

responses are transformed as information passes through the gustatory system. None of

the theories described above, including combinatorial coding, are specific enough to

describe the computations performed at any level of the gustatory system. The anatomy

suggests that gustation is complex and highly integrated with other systems in the brain.

A more specific theory of the system would describe why

26

Page 34: Gustatory Information Processing By Sam Reiter Thesis

1i. Taste receptors and receptor cells

Perhaps the most active area of recent research in gustation has been on the initial

representation of tastants. Here too has been the area of greatest contemporary debate. I

will first review some basic anatomy and physiology.

In vertebrates, tastants chemicals are initially detected when they contact the

tongue and soft palate. The tissues are studded with structures called taste buds, each of

which contains a collection of 50-100 tightly packed taste cells. Taste cells can be

divided into three classes, based upon the types of receptors and neurotransmitters that

they express88. The most prominent of these receptors are the taste receptors, of which

there are ~35 thus far identified in mice10.

Activation of receptors expressed in type two and three cells leads the generation

of all or none electrical action potentials89. Type one taste cells are thought to perform a

more supporting role, and do not spike.88 Most receptors are G-protein coupled, with

receptor activation producing a signaling cascade that elevates the internal calcium

concentration enough to cause the opening of TrpM5 channels90. The opening of these

channels allows sodium to enter the cell, depolarizes it, and causes action potentials.

Some other receptors work more directly: for example sodium activates a sodium

selective channel in type three taste cells, leading to sodium influx and depolarization91,92.

The details of how many chemicals that animals can behaviorally detect activate taste

cells remain unclear93,94.

It is hypothesized that the large depolarization of taste cells caused by action

potential generation is necessary for a variety of neurotransmitters to be released95. Type

two taste cells do not form synapses, instead releasing Adenosine triphosphate (ATP)

27

Page 35: Gustatory Information Processing By Sam Reiter Thesis

through pannexin hemichannels96 as well as the ion channel CALHM1.97 ATP excites

free afferent nerve fibers,98,99 type three taste cells, as well as the type two receptor cells

that released it100. Type three taste cells directly synapse onto afferent neurons,

presumably influencing their activity99. They also release serotonin (5-HT),101 which

inhibits type two taste cells, as well as self-inhibits type three taste cells100. Several other

neurotransmitters are expressed by taste cells and may play a role in gustatory

information processing. These include GABA102, norepinephorine103, acetylcholine104,

and glutamate105.

1st order afferent neurons with cell bodies in the geniculate ganglion project to

taste buds in the anterior of the tongue through the chorda tympani (CT) nerve, and to the

soft palate through the greater superficial petrosal nerve. The posterior tongue is

innervated by the glossopharyngeal nerve, with cell bodies located in the petrosal

ganglion10. The pattern of taste bud innervation by afferent neurons appears to be

somewhat species specific. In mice, 3-5 neurons target a single taste bud, while 2-16 have

been reported in rats, and 3-35 in hamsters. In mice, a single neurons appears to only

target a single taste bud, while in rats, cats, and sheep neurons have been shown to target

multiple buds106. While it is known that some neurons form synapses and some do not,

the specificity of taste cell-neuron connections is unknown. It is unclear, for example,

whether single neurons receive input from taste cells that express a certain combination

of gustatory receptors.

How is information about tastants represented and transformed as it passes

through these initial steps of the gustatory system? The first step to answering this

question is to describe the encoding properties of the taste cells. Work in this area has

28

Page 36: Gustatory Information Processing By Sam Reiter Thesis

been highly influenced by the idea of basic tastes. In particular, it plays a central role in

the research conducted by Zuker and Ryba. As their work has become the textbook view

on gustation107, it warrants close study. In their first words on the topic,

“Mammals taste many compounds but are believed to distinguish between only five basic taste modalities: sweet, bitter, sour, salty, and umami (the taste of monosodium glutamate). Although the discriminatory power of taste appears modest, it provides animals with valuable sensory information for the evaluation of food.”108

Zuker and Ryba differ from Pfaffman in their motivation for using BT

terminology. Pfaffman imported the BTs directly from human perception, without

argument. Zuker and Ryba associate the BTs with different proposed functions of the

gustatory system. This becomes clearer in a later review, where they explicitly assign a

small number of behaviors to each basic taste,

“Umami and sweet are “good” tastes that promote consumption of nutritive food (such as the building blocks for protein synthesis and energy), whereas bitter and sour are “bad” tastes that alert the organism to toxins and low pH, promoting rejection of foods containing harmful substances (for instance, noxious plants or spoiled or unripe fruits). Salt can taste either “good” or “bad” to us and be attractive or repulsive to mice, depending both on the concentration of sodium and on the physiological needs of the taster…”10

This link between the BTs and specific ecological functions serves as a very

potent motivation for finding how single BTs are initially represented by gustatory

receptors. If a receptor for a BT could be found, then through this link, a wide class of

behaviors could be explained. However, we have seen that gustatory behaviors cannot be

so simply categorized. Given the diversity of relationships different species have with

29

Page 37: Gustatory Information Processing By Sam Reiter Thesis

individual tastants, it seems unlikely that universal statements linking taste to behavior

can be accurately made. Despite this, Zuker and Ryba interpret the function of the ~35

gustatory receptors thus far discovered in terms of a single neuron code for the BTs. I

will next review both the evidence upon which this interpretation rests, and the

interpretation in detail.

1i.1. Bitter

The first gustatory receptors to be studied in detail were a family now recognized

to contain ~30 receptors, termed T2Rs108. Heterologous expression studies showed that

single receptors are sensitive to subsets of tastants considered bitter by humans. The

specificity of T2Rs appears to vary widely. Tested with 104 bitter chemicals, some

human T2Rs respond to only a single chemical, some to 50109. In mice the receptor T2R-

5 responds to cycloheximide, but not quinine, atropine, brucine, caffeic acid, denatonium,

epicatechin, phenyl thiocarbamide, 6-n-propylthiouracil, strychnine, or sucrose

octaacetate110. Mice lacking this receptor show a decreased (but not abolished) sensitivity

to cycloheximide, but normal sensitivity to the bitter tastant quinine111. This supports the

idea that an animal’s sensitivity to a broad range of tastants is accomplished by a large

number of receptors sensitive to different, overlapping subsets of tastants.

Consistent with this, polymorphisms in different T2Rs are associated with deficits

in detecting specific tastant chemicals110. Further, the large genetic sequence differences

between T2Rs across species can sometimes account for tastant specific differences in

detection ability. For example, humans describe b-glucopyranosides and

Phenylthiocarbamide (PTC) as bitter, but mice do not detect these chemicals. Mice

30

Page 38: Gustatory Information Processing By Sam Reiter Thesis

expressing human receptors sensitive to these chemicals in T2R expressing cells are able

to detect them111.

It is therefore clear that there is no bitter receptor. However, a case has been made

for bitter taste receptor cells (TRCs). Different T2Rs are co-expressed in the same taste

cells, suggesting that single TCs may be more broadly responsive to tastants than

individual T2Rs. To further test this idea, Meuller et al.111 removed a step in the signaling

cascade of T2Rs, PLC beta2. Mice lacking PLC beta2 lost sensitivity to several normally

aversive chemicals. PLC beta2 was then selectively expressed under the promoter of a

specific T2R. This rescued the aversive behavior in response to all the delivered aversive

tastants, not just the tastants that the specific T2R responds to. Identical results were

obtained when PLC beta2 was selectively restored during adulthood, addressing the

possibility that these results were due to developmental changes in the system. If the

selectivity for tastants seen in a heterologous expression studies matches the selectivity in

vivo, this is strong support for T2R expressing TRCs being more broadly responsive than

individual T2Rs111.

These findings have led to the hypothesis that by co-expressing most or all of the

T2Rs, certain TRCs function as general detectors of bitter taste.

“Our finding that each taste receptor cell expresses a large number of T2Rs is consistent with the observation that mammals are capable of recognizing a wide range of bitter substances, but not distinguishing between them.”108 “These results show that the bitter taste circuitry can be established without bitter sensory input, and unequivocally demonstrate that individual T2R cells operate as broadly tuned bitter sensors.”111

31

Page 39: Gustatory Information Processing By Sam Reiter Thesis

These conclusions reach past the experimental evidence, extending the inferred

sensitivity to 8 tested bitter compounds to the many thousands of chemicals that humans

taste as bitter, including but not limited to

“Hydroxyl fatty acids, fatty acids, peptides, amino acids, amines, amides, azacycloalkanes, N-heterocyclic compounds, ureas, thioureas, carbamides, esters, lactones, carbonyl compounds, phenols, crown ethers, terpenoids, secoiridoids, alkaloids, glycosides, flavonoids, steroids, halogenated or acetylated sugars, and metal ions.” 109

More direct measurements of TRC responses, done with calcium imaging from

TRCs in the tongue, reveal that different taste cells respond to different but overlapping

sets of bitter tastants112. Further, the tastant specific neural responses, animal behaviors,

and human perceptions discussed above cannot be explained if single TRCs function as

general detectors of bitter tastants.

1i.2. Sweet and umami

A second, smaller group of gustatory receptors are termed T1Rs. There are three

types of T1Rs, T1R1-3. T1Rs do not appear to be expressed in T2R expressing TRCs113.

T1R3 is sometimes co-expressed with T1R1, sometimes with T1R2, and sometimes with

neither113–115. In a heterologous expression system, mouse T1R2+3 expressing cells

respond to sucrose, fructose, saccharin, acesulfame-K, and dulcin, but not maltose,

glucose, galactose, aspartame, and several D-amino acids86,89,90. T1R1+3 expressing cells

respond broadly to L-amino acids116, and cells expressing T1R3 alone respond to high

concentrations of several sugars 117.

32

Page 40: Gustatory Information Processing By Sam Reiter Thesis

Mice lacking T1R2 or T1R3 have less sensitivity to sucrose, glucose, maltose,

and saccharine, although they are still attracted at high concentrations113. Mice lacking

both T1R2 and T1R3 do not respond to these chemicals117. Mice lacking T1R1 or T1R3

cannot detect several L-amino acids, and are impaired at detecting others117.

It has been argued that a dimer of T1R2+3 functions as the mammalian sweet

receptor, and by being expressed in cells not expressing T1R1 or T2Rs, make certain

TRCs selective for sweet taste.

“Indeed, functional expression studies in heterologous cells revealed that T1R3 combines with T1R2 (T1R2+3) to form a sweet taste receptor that responds to all classes of sweet tastants, including natural sugars, artificial sweeteners, d-amino acids and intensely sweet proteins.”118

Several lines of evidence call this assertion into question. T1R2+3 cells do not

respond to many things that humans taste as sweet, and that mice are attracted to. T1R3

responds to sugars, and is often expressed without T1R2 in a TRC. T1R2 knockout mice

are still attracted to many sugars, and display almost normal attractive behavior in

response to sweeteners like polycose and Aceflame K119–121. Further, certain chemicals

can inhibit the CT nerve responses to some sweet tastants and not others122, and single

afferent neurons can respond to different subsets of sugars in the same animal64. These

phenomena would not be possible if a single receptor or cell type mediated the receptivity

to sweet.

It has also been suggested that T1R1+3 serves as a receptor for umami taste.

33

Page 41: Gustatory Information Processing By Sam Reiter Thesis

“Final proof that T1R1+3 functions in vivo as the amino-acid (umami) taste receptor was obtained from the study of T1r1- and T1r3-knockout mice “118

The studies that this quote refers to, described above, strongly show that T1r1 and

T1r3 knockout mice display a deficit at detecting L-amino acids. There is some

controversy surrounding these knock out studies; a different T1R3 knockout mouse is

reported to largely retain the ability to detect MSG and sucrose123. Setting this issue

aside, the question becomes whether a broad amino acid receptor can be termed the

umami receptor. Zuker and Ryba, after studying the responses of T1R1+3 correctly

concluded

“…T1R1+3 responds to most L-amino acids, but not all amino acids taste the same: some are attractive to mice and sweet to humans, whereas others are neutral; some are even perceived as bitter and are aversive to animals.”116

Indeed, humans describe glycine, alanine, and threonine as sweet, proline as

sweet-acidic, monosodium glutamate and lysine as salty, arginine as salty-bitter,

phenylalanine as “obnoxious”, “repulsive” and “poisonous”, tryptophan as “ruinous”,

“minerally”, and “burning”, cysteine as “sulphurous” and “complex”, tyrosine as “weak”,

“flat”, and “dilute”, and leucine as “weak” “sweet”, or “tastless”29,78.

Mice and rats show preferences that roughly correlate with human perception of

L-amino acids, tending to prefer those described as sweet, avoid those described as bitter,

and be neutral to those humans describe as having a weak taste. Mouse preference is also

dependent on tastant concentration, with several L-amino acids attractive at low

concentrations and aversive at high. Interestingly, other animal species display

34

Page 42: Gustatory Information Processing By Sam Reiter Thesis

remarkably different preferences for L-amino acids. House cats, and the house musk

shrew (Suncus murinus) have preferences anti-correlated with those of mice124. These

behavioral results call into question the proposed simple links between amino acids,

umami taste, and “good” tastes. Because many amino acids that T1R1+3 respond to are

describes as sweet by humans and are preferred by mice, T1R2+3 cannot be the sole

sweet receptor. And because many amino acids that T1R1+3 respond to are describes as

bitter by humans and are aversive to mice, T2Rs cannot be the sole bitter receptors. Also,

because T2Rs and T1R1+3 are expressed in non-overlapping sets of TRCs113–115, cells

expressing multiple T2Rs cannot function as the sole bitter TRCs.

Rather than being the umami receptor, T1R3 is likely one of multiple receptors

that respond to L-amino acids. Calcium imaging of TRCs in the tongue revealed that

T1R3 knockout mice display physiological responses to many amino acids, including L-

glutamate125. A metabotropic glutamate receptor, mGluR-4, is expressed in TRCs,

responds to glutamate, and when knocked out produces a defect in detecting MSG126,127.

1i.3. Receptor addition experiments

One very interesting experiment deserves special mention. After establishing that

T1R2+3 responds to many chemicals mice are attracted to, and some T2Rs respond to

chemicals that mice are averse to, Zuker and Ryba expressed the synthetic opioid

receptor RASSL with either T1R2117, or with one of the T2Rs111. Control mice do not

detect the ligand for RASSL, spiradoline. In contrast, when RASSL is co-expressed with

T1R2, mice are attracted to the taste of spiradoline, and when co-expressed with one of

the T2Rs, mice avoid the taste of spiradoline. The authors conclude,

35

Page 43: Gustatory Information Processing By Sam Reiter Thesis

“Together, these results substantiate the coding of both sweet and bitter pathways by dedicated (that is, labelled) lines.” 111

Here ‘labelled lines’ refers to a single neuron code. By ‘pathways’, Zuker and Ryba mean

that single neuron (well, technically ‘cell’ rather than ‘neuron’, because TRCs are

epithelial cells) coding is maintained through a succession of neural populations, with

sweet TRCs selectively connecting with sweet 1st order neurons, connecting with sweet

2nd order neurons, etc. They argue that the fact that activating certain cells leads to

attractive behavior, and activating certain other cells leads to aversive behavior proves

that both these cells as well as cells further downstream can be cleanly associated with

different basic tastes (sweet and bitter) 111.

But, as previously discussed by several groups53,88, it is unclear how these

receptor addition experiments prove single neuron coding through a succession of neural

populations. To see why, Chaudhari and Roper offer a useful thought experiment:

“Take for example a computer keyboard. Striking the “A” key activates a combination of electronic signals that results in the illumination of a combination of pixels to produce the first letter of the alphabet on screen. If the plastic key (the “receptor”) on the keyboard were changed, striking the replacement key would still produce the letter “A” on screen. The experiment does not inform one about the electronic coding that is out of sight between the two visible events, and does not imply that labeled wires link the base of the key to particular pixels.”88

Chaudhari and Roper’s thought experiment reveals how the receptor addition

experiments do not constrain gustatory information processing. Certain TRCs respond to

tastants that mice are attracted to, and others respond to tastants that they are averse to.

36

Page 44: Gustatory Information Processing By Sam Reiter Thesis

The link between receptor activation and behavior is complicated and mysterious. The

receptor addition experiments activate these same TRCs using a novel stimulus. The fact

that this reproduces attractive or aversive behavior does not reveal anything about the

link between receptor activation and behavior.

1i.4. Sour

A nonselective polycystic kidney disease-like cation channel called PKD2L1 is

expressed in TRCs not expressing T1Rs or T2Rs. Killing these cells resulted in a loss of

CT nerve spiking in response to three acids presented to the tongue94. When co-expressed

in a heterologous expression system with another channel, PKD1L3, cells become

sensitive to acids128. These facts motivated the argument that a PKD2L1 receptor

complex is the receptor for the BT of sour, and by being expressed in unique population

of TRCs mice use a single neuron code.

“The results presented here establish that sour taste, much like our previous findings for sweet, umami and bitter taste, is mediated by a unique cell type, independent of all other taste qualities.” 94

These conclusions are inconsistent with the finding that mice lacking a functional

Pkd1L3 gene exhibit normal behavioral aversion in response to acidic tastants129. Other

channels have been proposed to partially mediate receptivity to acid tastants130.

1i.5. Salty

Using calcium imaging, it has been shown that NaCl and KCl excite different but

overlapping TRCs when applied to the tongue. The epithelial sodium channel (ENaC) is

37

Page 45: Gustatory Information Processing By Sam Reiter Thesis

expressed in TRCs, partially overlapping with PKD2L1 expressing cells. It responds

specifically to sodium and is active at low concentrations of NaCl application. The

integrated spiking activity of the CT nerve is normally very strong when NaCl is applied,

and relatively weaker when some other salts are applied. Knocking out ENaC selectively

reduces the CT nerve responses to NaCl application, making it approximately as strong as

the responses to some other salts. ENaC knockout mice display normal aversion to high

concentrations of NaCl and other salts, but do not show the wild type attraction to low

concentrations of NaCl91.

These findings were interpreted as support for single neuron coding of the BTs.

The BT of salt was partitioned, with ENaC functioning as a sodium detector and some

unknown receptors responding non-selectively to NaCl and other salts. Single neuron

coding was argued to apply because ENaC expressing cells were not co-expressed with

T1Rs and T2Rs. ENaC’s partial co-expression with PKD2L1 was deemed unimportant,

as

“Indeed there is total segregation of the cells responding to salt (low and high concentrations) versus those responding to acid stimulation (that is, sour cells never respond to salt stimuli….” 91

What receptors respond non-selectively to salts? Recently it has been shown that

The “bitter” T2Rs and the “sour” receptor PKD2L1 do. Removing either the PKD2L1

receptor or Trpm5, which is necessary for T2R function, reduces sensitivity to salts.

Rather than selectively abolishing aversion to bitter, sour and salty stimuli, mice created

with both receptors non-functioning did not display aversion behavior in response to a

38

Page 46: Gustatory Information Processing By Sam Reiter Thesis

wide range of delivered tastants, making it difficult to interpret the receptor’s combined

effects131.

These new finding are particularly interesting. They definitively show that

different salts are initially represented by different and overlapping sets of TRCs sensitive

to tastants associated with more than one BT and not all the tastants within a BT.

Calcium imaging studies show the same result132. Further, the complicated interactions

between taste cells result in making the taste cells that synapse onto 1st order afferent

neurons (type 3 cells) more broadly responsive to tastants than the TRCs133. This suggests

that rather than achieving their broad responsiveness by integrating the activity of

narrowly tuned TRCs, 1st order afferent neurons at least partially inherit their broad

responsiveness from taste cells.

1i.6. Summary

In sum, despite popular interpretation, close study of the initial representation of

tastes by TRCs calls into question the merits of describing the process using BT terms.

Single receptors do not respond to basic tastes. Instead, tastants within a basic taste are

recognized by multiple receptors, each only sensitive to some of the tastants associated

with a BT (T1R2+3, T1R3, ENaC) sensitive to multiple basic tastes (T1R1+3), or both

(T2Rs, PKD2L1). It is probable that we currently underestimate the degree to which

different receptors respond to overlapping sets of tastants, given the small numbers of

tastants used in most experiments. Multiple receptors are expressed in TRCs, but not in a

way that forms BT specific TRCs.

39

Page 47: Gustatory Information Processing By Sam Reiter Thesis

Besides drastically oversimplifying the diversity of receptor types, TRC types,

and gustatory behavior, at this level of analysis the BT interpretation artificially narrows

the types of stimuli that are considered to activate the gustatory system. There is a long

history of maintaining the BT interpretation by pushing the receptivity of certain

chemicals into other sensory systems. von Skramlik did so in 1926:

“As we have discussed, taste stimuli do not always cause pure taste sensations. Many testable substances also act on receptors adjacent to the taste organs. But various sensations can be produced by taste substances which have been proven to stimulate only the taste sense.” To von Skramlik, these “true” taste substances were of course those that could be

described in terms of the 4BTs. Here he comes dangerously close to a “no true Scotsman”

fallacy, with the 4BT partition of gustatory perception being maintained by designating

any exception to the partition as somehow not a taste.

Being specific about what constitutes a gustatory stimulus, and what should be

thought of as a stimulus for some other sensory system like olfaction, nociception,

thermoreception, or mechanoreception, is important. It can be seen as a step in the

process of determining the gustatory system’s stimulus space, which is necessary in order

to build accurate models of how neurons in the system respond to tastants. But the

separation of different stimuli into different sensory ‘systems’ is useful only if the

nervous system actually treats these stimuli differently. If it doesn’t, then our partition of

sensory space into different ‘systems’ stands on the same shaky ground as the BTs

themselves. Chaudhari and Roper define gustation in a less problematic way, stating

40

Page 48: Gustatory Information Processing By Sam Reiter Thesis

“Strictly speaking, gustation is the sensory modality generated when chemicals activate oral taste buds and transmit signals to a specific region of the brainstem (the rostral solitary nucleus).”88

This definition of gustation is pleasingly precise, making it clear how one would

classify some new stimulus as gustatory or not. It can be usefully modified by adding that

if some other nerve were discovered to convey information that is used to inform

perceptions or behaviors that are considered gustatory, it should be considered as part of

the gustatory system. After all, this link to perception and behavior is the only reason that

we associate nerves projecting to the rostral solitary nucleus with gustation in the first

place. Also, the stimuli activating the system do not strictly need to be chemicals. Zuker

and Ryba’s receptor addition experiments and Müller’s law of specific nerve energies

apply here: the perception of taste does not depend on the nature of what activates the

gustatory system, but how it becomes activated.

Under this definition, many stimuli have now been found that should be

considered gustatory, but are not universally recognized as such. One prominent example

is carbon dioxide (CO2). Humans can perceive the presentation of CO2. PKD2L1

expressing cells, which we now know to respond to acids and salts131, also respond to

CO2134. This seems to be accomplished through these TRCs expressing the carbonic

anhydrase Car4. Car4 knockout mice have weakened CT responses to CO2134.

Because CO2 directly activates TRCs, it should be considered a gustatory

stimulus. This proves problematic to the 5BT model, and especially a single neuron

coding model, as CO2 is distinct from the 5BTs, and activates ‘sour’ cells (which are also

activated by salts). Chandrashekar et al. dealt with this problem by arguing that because

41

Page 49: Gustatory Information Processing By Sam Reiter Thesis

CO2 activates non-gustatory sensory cells as well, it can be considered as a multisensory,

rather than a taste stimulus.

“Humans perceive five qualitatively distinct taste qualities: bitter, sweet, salty, sour, and umami (a savory sensation characterized by the taste of monosodium glutamate). Sweet and umami are sensed by members of the T1R family of heterotrimeric guanine nucleotide binding protein (G protein)–coupled receptors (GPCRs) (1–3); bitter stimuli are detected by T2R GPCRs (4–7); and sourness is sensed by cells expressing the ion channel PKD2L1 (8–10). In the tongue, these receptors function in distinct classes of taste cells, each tuned to a specific modality (7, 8, 11, 12). In addition to these well-known stimuli, the taste system appears to be responsive to CO2” 134 “Although CO2 activates the sour-sensing cells, it does not simply taste sour to humans. CO2 (like acid) acts not only on the taste system but also in other orosensory pathways, including robust stimulation of the somatosensory system (17, 22); thus, the final percept of carbonation is likely to be a combination of multiple sensory inputs.” 134

There are many more examples of stimuli not recognized as proper tastants that

activate the same cells as the 5BTs. Mice and humans can discriminate the taste of

calcium, a chemical vital for many cellular functions135. High concentrations of calcium

are avoided by mice unless deprived of calcium in their diet. The “sweet” receptor T1R3

responds to the presentation of calcium. It has been shown to have the same calcium

binding pocket residues as the calcium sensing receptor. T1R3 KO mice do not avoid

high concentrations of calcium.136 Washing some tastants out of the mouth with water

produces a ‘water taste’. This is caused by some tastants, including the artificial

sweetener acesulfame-K, actually inhibiting T1R3. Washing with water afterwards

disinhibits the receptor, producing the perception of a taste137. The temperature sensitivity

of TRPM5, a channel downstream of T1R and T2Rs in gustatory signal transduction,

serves to make the strength of TRC activation dependent on local temperature138. This

may underlie the fact that humans perceive a taste when portions of the tongue are

42

Page 50: Gustatory Information Processing By Sam Reiter Thesis

warmed or cooled139.The receptors GPR40 and GPR140 respond to the presentation of

fats. These receptors are expressed in TRPM5 expressing TRCs. In a 2 bottle choice test

mice prefer the taste of fatty acids, and KO mice for either receptor lowers their

preference, as well as the integrated CT nerve response140. Capsaicin is the chemical

found in chili peppers that feels hot and burning when eaten. The capsaicin receptor

VR1141 is expressed in most TRCs expressing T1R2 “sweet” or T1R3 “sweet” and

“umami”, and in ~96% of T2R6 “bitter” expressing TRCs142. TRPM8, a receptor

responding to cold temperatures and the chemical menthol143 is expressed in nerve fibers

innervating taste buds144.

Narrowing the scope of gustation to align with how the nervous system initially

represents different types of stimuli is important. However, the above examples show that

this has been taken too far. There doesn’t appear to be any principled reason that the

many kinds of stimuli that activate the same taste cells as traditional tastants are not

considered to be legitimate tastes. This is problematic for the BT model, and single

neuron coding of BTs in particular. However, with these different stimuli activating

different but overlapping sets of cells, unique representation is possible under a

combinatorial code.

1j. Insect Studies

Constituting over 80% of all animal species, insects deserve to be studied in their

own right. But with relatively simple accessible nervous systems, insects offer the

opportunity to uncover basic principles about how nervous systems work in general. The

43

Page 51: Gustatory Information Processing By Sam Reiter Thesis

physiological complexity of the mammalian gustatory system makes insects especially

attractive models for studying the neural basis of gustation.

In insects, tastant chemicals are initially detected when they come into contact

with specialized hair cells called sensilla that stud the antenna, mouthparts, wings, legs,

and abdomen of most insects145. Each sensillum houses two to four neurons, a mixture of

gustatory receptor neurons(GRNs) and mechanosensory neurons146. GRNs in the legs and

wings project to ganglia in the insect body, while GRNs in the mouthparts and antennae

project to a region of the insect brain called the sub-esophageal zone (SEZ). Each GRN

expresses several receptors for tastants. The most common of these receptors are termed

gustatory receptors (GRs). The number of GR types expressed depends on the insect

species. 68 have been identified in Drosophila melanogaster, 114 in Aedes aegypti

mosquitos, 10 in the honeybee Apis mellifera, 17-97 in different ant species, and 58 in

the water flea Daphnia pulex147.

The expression pattern and intracellular signaling cascade of GRs has not been

fully characterized147, but it is clear that some receptors are expressed in non-overlapping

sets of neurons, some are expressed in completely overlapping sets of neurons, and some

are partially overlapping147.

Since the 1950’s, the spiking activity of GRNs has usually been studied using the

tip recording technique. A tastant is loaded into a glass pipet which is then fitted over

one of the animal’s sensilla148. Although this approach allows for an extracellular

recording of the responses of GRNs, the tip technique has several limitations. Assigning

the extracellular responses to a single GRN is difficult because there are several GRNs in

each sensillum, and delivering a range of tastants requires making many sequential tip

44

Page 52: Gustatory Information Processing By Sam Reiter Thesis

recordings. Also, every tastant must be mixed with a conductive salt solution to allow

recording the neurons’ electrical activity. Further, potentially important properties of the

response dynamics of GRNs cannot be recorded using this technique; it is impossible, for

example, to know how a GRN responds immediately after a tastant has been removed.

Finally, by itself, this technique does not allow recording the responses of higher order

gustatory neurons.

Classic work using this technique, extensively reviewed by Detheir146, concluded

that the responses of GRNs could be classified into just a handful of types, with

sensitivities for water, sweetness, saltiness, and bitterness. More recently, however, it was

found using the tip technique that fly GRNs exhibit a more diverse range of response

profiles. In this respect strikingly similar to mammals, many GRNs responded to

chemically diverse tastants, and with a variety of temporal activity patterns. Further, a

given tastant could elicit responses from a population of different GRN types, and these

neurons responded with tastant-specific changes in spike rate147,149. It seems possible that

GRNs exhibit more complicated temporal patterns of spiking activity that cannot be

assessed comprehensively using the tip technique.

GCaMP imaging has also been used to record the responses of GRNs.150 By co-

expressing GCaMP along with different GRs it is possible to measure calcium activity in

GRNs expressing a specific GR. This is done through imaging the axons of many GRNs

in the SEZ. Experiments using this approach suggest that different GRNs respond to

different and largely non-overlapping groups of tastants10,150,151. However GCaMP

imaging of GRN axons has several rather severe limitations. The relationship between

GCaMP signal and spiking activity is uncharacterized, but most likely nonlinear152,

45

Page 53: Gustatory Information Processing By Sam Reiter Thesis

making it difficult to interpret the meaning of a signal. And because different GRs are

expressed combinatorially within a single GRN153, co-expression of GCaMP with a

single GR actually expresses GCaMP protein in a heterogeneous population of GRNs.

Thus the presence of a GCaMP response cannot indicate that a specific GRN type is

activated, or how it is activated. Like the tip recording technique, GCaMP measures

from GRNs fall short of ideal in ways that compromise our ability to understand how

these receptor neurons function.

Thus, while insects represent attractive models in which to study fundamental

questions about gustatory information processing, existing methods do not allow for

many of the questions debated in mammalian gustation to be addressed.

46

Page 54: Gustatory Information Processing By Sam Reiter Thesis

2. RESULTS

2a. Moth gustatory behavior

The gustatory system is responsible for detecting the thousands of chemicals that

animals contact, and informing many important behaviors. It has been suggested that this

is accomplished relatively simply, where every chemical is associated with a small

number of ‘basic tastes’10,17,47. It is debated whether different basic tastes are initially

represented using non-overlapping subsets of cells (labeled lines)87,94,111,116,117,150, or

overlapping subsets (across fiber patterns)53,59. Testing these hypotheses requires detailed

knowledge of how tastes are initially represented within the population of receptor cells,

and how this representation is transformed as it moves to higher order populations. This

has proven difficult given the system’s physiological complexity and the technical

difficulty of delivering tastant chemicals while recording cellular activity in vivo.

Insect models offer the potential for such a point to point analysis154. Here I

studied the gustatory system of the adult moth Manduca sexta. I found that moths, like

other animals, could produce tastant specific behaviors. Recording the tastant induced

activity of first and second order gustatory neurons, I found that individual tastants are

uniquely represented as spatiotemporal patterns of activity distributed across the GRN

population. Tastant representations are then substantially transformed by postsynaptic

neurons. My results suggest that in order to generate chemical specific behavior,

chemical specific information is processed by the gustatory system.

Insects can use gustatory receptor neurons (GRNs) located in sensilla studding the

distal 1/3 of the proboscis, mouthparts modified to form a drinking straw like organ, for

47

Page 55: Gustatory Information Processing By Sam Reiter Thesis

gustatory discrimination and conditioning146. When a high concentration of sucrose is

applied to the proboscis, Manduca are known to reflexively extend their proboscis155.

Most animals are capable of producing different behavior in response to different

tastants 36,81,156,157. I sought to find tastants that Manduca react to differently. Any two

tastants that produce different behavior must be distinctly represented by the nervous

system, following the data processing inequality4. Finding such tastants would thus

provide a useful constraint on the interpretation of neural activity. Working with the

assistance of the high school intern Chelsea Campillo, I tested the tastant specificity of

Manduca proboscis extension behavior by delivering a range of tastants to the proboscis.

Behavioral and physiological experiments were performed on moths reared from

eggs on an artificial diet4 at 26 degrees in >70% humidity. We tested the probability of

proboscis extension in response to a variety of stimuli. 1-2 day old moths were vertically

fixed in tubes with their head exposed and the most proximal 1/3 of the proboscis secured

in a small tube.5 Moth’s eyes were colored with ink to eliminate visual cues. A behavioral

test consisted of delivering 50 µl of a stimulus being delivered to the distal 2/3 of the

proboscis. The subsequent 5 minutes of behavior was recorded by a digital camera (Flea

3, Point Grey). After this, the proboscis was washed with distilled water prior to the next

test. A single moth was used for 5 behavioral tests, each separated by 5 minutes. Epochs

of proboscis extension were manually scored from video using a GUI custom written in

Matlab (Mathworks). Moths were considered to have extended their proboscis if they

raised it for >5 seconds following tastant presentation. Tastants were presented in a

random order, and the experimenter was blind to tastant identity throughout testing and

analysis.

48

Page 56: Gustatory Information Processing By Sam Reiter Thesis

We chose chemically diverse tastants commonly used in gustatory behavioral

experiments. For all tastants except caffeine and lobeline, equal proportions of 50, 250,

500, 750, and 1000 mM concentrations were delivered pseudo randomly. Caffeine and

lobeline do not dissolve into water at these concentrations, so caffeine was presented at

1/10, and lobeline at 1/100 these concentrations, as done previously149. For each tastant,

results were pooled across all concentrations to show concentration independent

differences in tastant behavior (Fig. 1).

All the presented tastants caused moths to extend their proboscis above the rate in

response to distilled water alone. The rate of proboscis extension varied with tastant

identity; the presentation of sucrose elicited much more frequent extension than all other

presented tastants (Chi squared test with Bonferroni correction, p<0.05). This included

the other tested sugars, trehalose and maltose. Although moths sometimes extended their

proboscis in response to these other sugars, they extended their proboscis just as often to

NaCl, KAc, caffeine, and lobeline (Chi squared test, p>0.05 after Bonfferoni correction).

Thus, differential proboscis extension behavior shows that moths can recognize the

presence of diverse chemicals, and can discriminate the taste of sucrose from that of two

other sugars.

49

Page 57: Gustatory Information Processing By Sam Reiter Thesis

Figure 1: Tastant specific moth behavior. The rate of proboscis extension following the presentation of the listed tastants (n=89.75 +/- 0.25 moths/tastant). Red error bars are s.e.m.

2b. A new method for studying the gustatory system

The moth’s nervous system could use a variety of strategies to represent tastants

that would be consistent with the behavior we observed. I sought to understand which

strategy is used by recording the activity of GRNs. I first designed a method to deliver

tastants at high temporal precision while recording neural activity.

The moth’s proboscis was fixed with epoxy into a rigid tube. A small piece of

mesh was installed 5 mm above the proboscis. The proboscis coil rested against this

mesh, ensuring that the proboscis was placed in the same location across moths. Filtered

water was pumped over the proboscis using a peristaltic pump (Manostat compulab 3) at

40 ml/min. The output tube of a pressurized 16-valve perfusion system (Bioscience

50

Page 58: Gustatory Information Processing By Sam Reiter Thesis

Tools) was installed 1 cm above the mesh. To deliver a tastant, pressure was applied with

a Pico Pump (World Precision Instruments). This was followed by the opening of a single

valve, injecting tastant into the water stream. To record the precise timing of tastant

delivery, I colored tastant solutions with a small amount of food coloring (Fast green

FCF, Sigma. 0.05% w/v). The tastant concentration present over the proboscis was

monitored with a color sensor (FS-V31, Keyence) installed 5 mm below the perfusion

system output, just above the mesh. Color sensor signals were amplified 5x by a DC

amplifier (Brown-Lee Model 440). The delivery system was controlled and color sensor

recorded using custom designed Labview software (PCI-MIO-16E-4 DAQ cards,

National Instruments).

Figure 2: A schematic of the tastant delivery system.

51

Page 59: Gustatory Information Processing By Sam Reiter Thesis

To estimate the concentration of tastant that reaches the moth after dilution in the

water stream, I injected tastants into water colored with the same concentration of dye as

the tastants. This was taken as 100% concentration delivery. I then observed what

fraction of color sensor signal was generated by injecting colored tastant into uncolored

water. In this way, I estimated that after dilution in the water stream the concentration

that reaches the moth is ~77% that of concentrations listed throughout the text (Fig. 3).

Figure 3: Estimating the concentration of tastant that reaches the moth. (Blue) The color sensor voltage trace in response to tastant delivery into uncolored water. (Red) as blue trace except the uncolored water has been colored with the same concentration of dye as the tastants. Traces = mean of 5 trials of each condition.

52

Page 60: Gustatory Information Processing By Sam Reiter Thesis

Using color sensor intensity as proportional to tastant concentration is valid

because the sensor’s responses were found to be linear throughout the range of color

intensity used in this, and all subsequent experiments(Fig. 4).

Figure 4: Color sensor reading is proportional to tastant concentration. Diluting colored tastant by the amounts listed on the Y axis produced a linear decrease in color sensor voltage. (Blue Dots) Single trial color sensor amplitude averaged over 1s tastant delivery. (Red) Best fit line.

To record from GRNs while tastants were delivered in this way, intact moths were

fitted into tubes with their heads exposed. After protecting the proboscis and antennae

with small plastic tubes secured with epoxy, a wax cup was built up surrounding the

ventral side of the head capsule. The dorsal part of the head capsule was first removed

and the Buccal Compressor158 muscle (a large muscle used for drinking fluids) was cut

for recording stability. This opening was then sealed with wax. The ventral side of the

head capsule was then bathed in Manduca physiological saline159. The Labial Palps,

exoskeleton, and trachea covering the brain were removed, exposing the maxillary nerve.

53

Page 61: Gustatory Information Processing By Sam Reiter Thesis

Saline was then removed and replaced with crystals of Collagenase/Dispase (Roche

Diagnostics) dissolved in saline (10% w/v). After this caused the Dilator-3158 muscles to

tear (large muscles used for proboscis extension, 2-3 minutes), the collagenase mixture

was replaced with saline. The sheath covering the brain was then removed. For stability,

a platform made from a flattened insect pin was inserted through the esophagus

underneath the SEZ, and secured to the tube with wax. During experiments, a constant

flow of fresh saline was passed over the brain.

The GRNs of the proboscis project into the brain through the maxillary nerve.

The maxillary nerve in Manduca is purely afferent, containing GRNs and

mechanosensory neurons. My tastant delivery system was designed in part to minimize

the change in mechanosensory input to the proboscis, in order to isolate gustatory

responses. Motor neurons evolved in proboscis movement project through other

nerves158. I was able to make sharp intracellular recordings selectively from the axons of

GRNs by targeting the nerve.

Figure 5: The maxillary nerve. Shown is a frontal view of the moth brain (z stack), with the GRNs of the left Maxillary nerve (indicated with dashed ellipse) filled with Rhodamine-Dextran projecting mostly into the left half of the SEZ.

S

Esophagus

SEZ

54

Page 62: Gustatory Information Processing By Sam Reiter Thesis

My intracellular recordings revealed that different GRNs responded to different

tastants, with different changes in spike rate. Some of these responses were locked to the

timing of the stimulus, and were quite reproducible over many trials (green box in Fig. 6).

Other tastant responses outlasted the stimulus by many seconds (ex. GRN 4 in Fig. 6).

Both excitatory and inhibitory responses were observed. Also, compared to receptor cells

in other sensory systems, GRNs exhibited remarkably little response adaptation, even

when activated very strongly by a prolonged stimulus.

Figure 6: GRN intracellular recordings. Raster plots of 4 example GRNs, presented with several tastants and concentrations. Multiple trials of the same tastant (5-7) are shown. (Left, Green box) A closer view of GRN 1’s responses to 1000 mM sucrose. (Bottom, blue) Intracellular voltage trace of the last trial in each raster where there was an action potential. (Bottom, red) Color sensor voltage, showing tastant delivery time (1s.).

The instability caused by the small size and flexibility of the maxillary nerve

made holding intracellular recordings for more than 5-10 minutes impossible. In order to

increase the number of GRNs that I could record from, as well as the number of tastants

that I could deliver during a single experiment, I developed an extracellular recording

technique.

55

Page 63: Gustatory Information Processing By Sam Reiter Thesis

When I placed bundles of handmade twisted wire extracellular electrodes

‘tetrodes’ into the maxillary nerve, I could record very large extracellular signals from

several GRNs simultaneously. In order to extract the activity of single units (neurons), a

method known as spike sorting needs to be applied. My laboratory’s method of spike

sorting160 was not able to be easily modified to deal with the wide distribution in spike

sizes or the extremely narrow spike waveforms that GRNs produced. It further did not

allow for a quantitative assessment of errors in sorting. I therefore worked with my

colleagues Zane Aldworth, Nitin Gupta, and Mark Stopfer to develop a novel computer

assisted spike sorting method. We combined the desirable features of Christoph Pouzat’s

Spikeomatic160 and the Klienfeld laboratory’s UltraMegaSort2000161,162 into an easily

modifiable set of Matlab scripts. This took the form of using Pouzat’s method of dealing

with waveforms overlapping in time (superpositions) and Klienfeld laboratory’s method

of estimating the percent of sorting errors. Our method is presented in detail in an

appendix (section 4).

2c. Profiling the responses of GRNs to tastants

Using my tastant delivery system and extracellular recording technique, I

recorded the activity of 83 GRNs in response to a 1s square wave of 12 tastants (Fig. 7).

Tastants and concentrations were selected based on either known ecological importance

to moths34, or because of association with one of the basic tastes in previous

literature149,150. Tastants used were sucrose, glucose, trehalose, maltose, sodium chloride,

lithium chloride, and potassium acetate (1000 mM), caffeine (100 mM), and (-)-lobeline

56

Page 64: Gustatory Information Processing By Sam Reiter Thesis

hydrochloride, berberine chloride, and denatonium benzoate (10 mM), with filtered H2O

delivered as a control.

I observed a great diversity in GRN resting activity (ex. GRN 1, 67), in the set of

tastants that caused different GRNs to respond (ex. 1, 31, 77), and the temporal structure

of these responses (ex. 27, 48, 67). For example, GRN 43 had a low baseline spike rate of

0.3 Hz, responded to the presence of NaCl, LiCl, and caffeine with bursts of spikes of

different peak rates (32.11, 67.91 and 22.19 Hz respectively), responded to KAc with

spiking that outlasted tastant stimulation, and responded to lobeline and berberine with

spiking activity that only peaked after tastant removal (40.13 and 52.21 Hz) (Fig. 8).

Figure 7: GRNs exhibit a diversity of tastant responses. Raster plots of 83 GRNs (rows) responding to 12 tastant stimuli (columns). Tastant presentation time is marked with shading. Scale bar is 5 seconds. Rows are ordered by hand, based roughly on correlation across tastant responses.

57

Page 65: Gustatory Information Processing By Sam Reiter Thesis

Figure 8: The temporally structured responses of Fig. 7’s GRN 43. The trial average spiking responses were smoothed with a Gaussian filter (s.d. 60 ms.). (Bottom, Red) Average color sensor voltage trace showing tastant delivery time.

GRNs exhibited a variety of tuning widths, with some GRNs responding to only a

single tastant, and others responding to most tastants delivered (Fig. 9a). Because the

notion of ‘response’ depends on how sensitive populations of unknown downstream

neurons are to spiking activity in GRNs, I calculated this tuning width for a large number

of thresholds. This was done as follows:

Beginning with multiple trials of a neuron responding to a tastant, I first estimated

the firing rate of the neuron by smoothing the trial averaged spike times with a Gaussian

filter (s.d. = 60 ms.). I defined the background activity as the firing rate of the neuron in

the first second of the trial, before tastant delivery. A neuron was considered to respond

to a tastant if its firing rate exceeded the mean background activity + N standard

deviations of the background activity, or fell below the mean background activity - N

58

Page 66: Gustatory Information Processing By Sam Reiter Thesis

standard deviations of the background activity. N refers to a variable threshold. I tested a

wide range of thresholds (1-12), moving from well below to well above generally

accepted response values163 to demonstrate the robustness of the results.

Notably, sucrose activated significantly more neurons than other tastants (at an 8

s.d. threshold, Chi squared tests, p<0.05 after Bonfferoni correction). ~50% of GRNs

were excited by sucrose (Fig. 9b, 54.22% for 8 s.d. threshold), and ~50% of neurons that

responded to sucrose did not respond to any other delivered tastant (Fig. 9c, 53.33% for 8

s.d. threshold). Other tastants activated ~25% (24.34+/- 0.77% for 8 s.d. threshold) of the

GRN population, mostly GRNs that responded to multiple tastants.

Figure 9: Summary statistics of GRN population data. All plots show results over response thresholds 1:12. (A) The percent of the GRN population that responded to varying numbers of tastants. (B) The percent of the GRN population that responded to the listed tastant. (C) The percent of neurons that responded to the listed tastant and did not respond to any other delivered tastant.

I next studied the concentration dependence of GRN activity. I recorded from

GRNs while delivering 4 second square pulses of varying concentrations of sucrose (Fig.

10). I delivered H2O and NaCl as control tastants.

C B A

59

Page 67: Gustatory Information Processing By Sam Reiter Thesis

Figure 10: The tastant responses of GRNs are concentration specific. Raster plots of 3 simultaneously recorded GRNs responding to multiple tastant concentrations. (Bottom, red) Average color sensor voltage trace signaling tastant delivery.

Most GRN activity depended on tastant concentration in a simple and reliable

way: increasing concentrations activated increasing numbers of GRNs with stronger

excitatory or inhibitory responses. (Fig. 11, 1 way ANOVA, F(4, 95) = 9.72, p<1e-5, post

hoc Tukey’s HSD, p<0.05). ‘Response strength’ was defined as the absolute value of the

difference between the mean firing rate during the stimulus time and the background

activity.

Figure 11: GRNs respond to increasing concentration with increased response strength. The mean response magnitude of a population of GRNs (n=20) responding to the concentrations of sucrose listed in Fig 10. Blue lines/red circles are the responses of individual GRNs, black line is the population mean +/- s.e.m.

60

Page 68: Gustatory Information Processing By Sam Reiter Thesis

I confirmed that this relationship held for a variety of tastants, using data collected from a

larger set of GRNs responding to 1s presentations of tastant (Fig. 12, 1 tailed Wilcoxon

signed rank tests, p<0.05).

Figure 12: The concentration dependency of GRNs is consistent across tastants. The strength of GRN responses to two concentrations of sucrose (n=75), NaCl (n=51) and caffeine (n=44). Lines connect circles representing a single GRN’s responses. (Blue bars) Mean response strength. For display purposes, data were normalized by dividing all response strengths by the strongest response to that tastant.

2d. Describing the responses of GRNs over time

I next sought to better describe the observed diversity in the temporal dynamics of

GRN activity. Traditionally this has been done qualitatively, with descriptions of how a

neuron’s trial averaged spike rate depends on the nature and timing of a sensory stimulus.

Relatively recently, variants of a statistical method called Generalized Linear Modeling,

(GLMs), or Linear-Nonlinear Modeling (LNMs) 164–166 have been usefully applied to

quantitatively describe how neurons in the visual167–169, auditory170, somatosensory171,

and olfactory172,173 systems depend on the temporal structure of stimuli. These two terms

name the same class of model, and are a historical legacy. It seems that two separate

communities of scientists began using these models, and only joined after independently

61

Page 69: Gustatory Information Processing By Sam Reiter Thesis

developing two different nomenclatures. I will use the LNM terms in this manuscript

because they are more neuroscience specific.

This modeling approach offers many advantages. It can be estimated using single

trials, and so doesn’t rely on the often dubious assumption of statistical stationarity or the

arbitrary choice of binning/smoothing parameters that underlies spike rate estimates.

Once fit, the model provides a prediction of how a neuron will respond to future stimuli.

If the model prediction does a good job of explaining neural activity, than we can say

with some confidence that we actually understand the transformation from stimulus to

spiking at the ‘algorithmic level’174. If the prediction fails, we can use the specific ways

in which it fails to motivate further hypotheses about what is being encoded by the

neuron’s spiking activity. And if a new model is constructed, we can quantitatively assess

the improvement over alternative models. Traditional methods offer little to no predictive

power. If after observing a neuron’s response to a certain stimulus, one would like to

predict the responses to a large class of other stimuli, a generative model like the LNM is

needed.

Perhaps even more importantly, LNMs are able to incorporate covariates166. In

most contexts, the spiking activity of a neuron depends on a large number of factors. For

example, the activity of a retinal ganglion cell will depend in some way on recently

presented visual stimuli, the neuron’s spike history, the spiking activity of other neurons

in the retina, the state of adaptation of receptor cells, etc. Traditional methods deal with

covariates like these by attempting to vary one factor at a time while holding everything

else constant. LNMs and related models offer the ability to measure many covariates at

62

Page 70: Gustatory Information Processing By Sam Reiter Thesis

the same time, and provide an estimate of their relative importance. This allows one to

see how all the ‘pieces of the puzzle’ fit together in a quantitative way.

As the name suggests, a LNM is a generalization of linear regression. My specific

goal was to model the relationship between the time varying concentration of a tastant

and the spike times of a GRN. I did so by modeling the spike times as draws from an

inhomogeneous Poisson process, that is, a time series where the probability of spiking at

any time is completely specified by a single spike rate parameter. I could then relate the

temporally varying stimulus to this estimate of the spike rate as a linear transformation

followed by a static, “spiking” nonlinearity. This nonlinearity performs the important role

of preventing the model from predicting pathological things like negative spike rates, and

allows for a greater range of stimulus-response relationships to be captured.

Mathematically the model is simply

P(spike/t) = F(k*s(t))

Where s(t) is the stimulus over some period t, k is a linear transformation

(“filter”) of the stimulus, F is a static nonlinear transformation, and the output of this is a

prediction of the probability of spiking during that time, P(spike/t).

For the curious reader, covariates can be incorporated into this model simply by

adding extra linear terms. So I could model how a neuron’s spike rate depends on a

stimulus, s, past spike history h, and the spike history of a number of simultaneously

recorded neurons n by the model

P(spike/t) = F(k*s(t) + j*h(t) + l*n(t))

63

Page 71: Gustatory Information Processing By Sam Reiter Thesis

These type of models are fit by finding the k (and j, l, and linear terms associated

with any other covariates) and the parameters of the chosen nonlinearity that best relates

the two sides of the equation169. This can be done by maximizing the likelihood of the

model parameters. In the case of linear regression there is a closed form maximum

likelihood solution, but in the LNM case no such solution exists. Instead, I initialize the

model with some guess for all the parameters, and then estimate how the likelihood

changes in the local region around the parameters. The parameters are then moved in that

direction, and I take a new estimate of how the likelihood changes locally. In this way the

likelihood is iteratively maximized. One can imagine this process as pushing a ball (the

model parameters) up a hill (the likelihood surface) in the direction where the hill is

steepest. When parameters climb to the top of the hill, I stop iterating and have the model

fit.

This process runs into trouble when the choice of nonlinear function makes the

likelihood surface non convex, that is, when there is more than one hill. In this case

iterative improvements of the likelihood could end on a local maximum that is less likely

than the global maximum likelihood solution. That is, I push the ball up the nearest hill

and stop, but there is a taller hill out there. I avoided this problem by choosing a

nonlinearity of a form that guarantees the convexity of the model’s likelihood surface175.

Fitting models by maximizing their likelihood allows for educated guesses about

the shape of the stimulus-response relationship to be incorporated. For example, one

suspects that if a neuron is affected by the stimulus history in the previous 100 ms., than a

1 ms. stimulus 54 ms. previous and a 1 ms. stimulus 53 ms. previous should elicit very

64

Page 72: Gustatory Information Processing By Sam Reiter Thesis

similar responses. Another way of saying this is that one can expect the temporal

receptive field of a neuron to be smooth. This expectation can be incorporated into my

model fitting procedure, where I penalize the likelihood of models with parameters that

are not smooth. This process is called model regularization75, and is a very useful way of

increasing the predictive power of LNMs by preventing over fitting to training data.

To describe how GRNs depended on the time course of tastant stimulation, I

delivered a wide range of temporal stimulus patterns of tastant concentration, while

recording extracellularly from GRNs as above. The temporally rich stimulus (1000 mM

sucrose, 27/35 experiments, or 2000 mM NaCl, 8/35 experiments) was constructed by a

pseudorandom sequence of Pico pump and perfusion system valve opening times. The

Pico pump was opened 800 times in a 7.5 minute trial. Opening times were drawn from a

uniform distribution spanning the trial length. Opening durations were drawn from a

uniform distribution spanning 8-608 milliseconds. Valves were opened 1600 times in a

7.5 minute trial. Opening times were drawn from a uniform distribution spanning the trial

length. Opening durations were drawn from a uniform distribution spanning 8-458

milliseconds. Stimulus parameters were selected to deliver a range of intensities and

temporal frequencies. Similar modeling results (see below) were generated using stimuli

of different lengths (5-10 minutes), and different opening durations (4-54 milliseconds).

My tastant delivery system allowed for the reliable presentation of the same temporally

rich stimulus over many trials (Fig. 13).

65

Page 73: Gustatory Information Processing By Sam Reiter Thesis

Figure 13: Trial to trial reliability of temporally rich tastant delivery. Three superimposed trials of the color sensor voltage trace. This color sensor signal is proportional to tastant concentration, given the linearity of the color sensor (Fig. 4)

With the assistance of my colleague James McFarland, I used his NIM toolbox168

(http://www.clfs.umd.edu/biology/ntlab/NTlab/NIM.html ) for fitting linear-nonlinear

models (LNMs) models of GRN activity. I used the function

F[x; a,b,h] = a*log[1 + exp(b(x-h))]

as the LNMs’ spiking nonlinearity, where x is the linear prediction of the model, and a, b,

and h are parameters of the nonlinearity. LNM linear filters and nonlinearity parameters

were fit by maximizing the regularized log likelihood. The likelihood was regularized

66

Page 74: Gustatory Information Processing By Sam Reiter Thesis

with a smoothness penalty (L2 norm), with smoothness parameter values picked to

maximize the cross validated maximum log likelihood. Models were fit for all recorded

GRNs that produced more than 5 spikes during each of the 5 test sections of data

(amounting to a minimal spike rate of >0.0625 Hz).

Model quality was assessed by a likelihood ratio test comparing the cross

validated log likelihood (LLx) of the LNM to that of the ‘null model’, which assumes a

constant spiking rate equal to the mean spike rate of the GRN throughout the test data.

This likelihood ratio was then normalized by the number of spikes in the test data to

allow for comparison of model quality across GRNs (bits/spike). The likelihood ratio is

an attractive measure of model quality because it can be estimated on single trials, and

provides a direct measure in the average reduction of uncertainty about the stimulus that

the presence of a spike provides (Shannon information)167.

GRN spike rate was quite well described by LNMs which related tastant

concentration to GRN spiking as a linear transformation of a time-embedded

representation of tastant concentration (temporal filter) passed through a static spiking

nonlinearity. While we are able to quantitatively assess the model quality based on a

single trial, to build intuition about the LNM’s goodness of fit I performed an additional

test on one GRN167. After fitting a model as above, I delivered 32 repeats of a 20 second

long stimulus with the same statistics as the longer nonrepeating stimuli while recording

GRN activity. The trial averaged spiking response was smoothed with a Gaussian filter

(s.d. = 60 ms.). This was compared to the LNM spike rate prediction from a single trial.

For a visual comparison with the GRN raster plot, 32 trials were simulated by Poisson

draws from the LNM spike rate prediction. (Fig. 14)

67

Page 75: Gustatory Information Processing By Sam Reiter Thesis

Figure 14: Comparing the rate of a GRN to that predicted under a LNM. (Top, Blue) 32 trials of a GRN responding to a temporally rich stimulus. (Black) 32 trials simulated from a LNM previously fit to the GRN’s activity. (Bottom, Blue) The PSTH of the GRN. (Black) The spike rate prediction of the LNM. (Red) the color sensor voltage trace, showing stimulus delivery (sucrose).

We can see from Fig. 13 that the LNM does a very good job of predicting the

GRN’s response to different concentrations and temporal patterns of sucrose. Indeed,

looking at the 32 trials of spiking, it is difficult to tell which were generated from the

GRN and which from the LNM. For this model, the LLx improvement over the null

model was 0.59 bits/spike.

68

Page 76: Gustatory Information Processing By Sam Reiter Thesis

The mean LLx improvement for all the models was 0.49 +/- 0.05 bits/spike. As I

modeled even largely silent GRNs, I expected to observe a wide range of likelihood

improvements over the null model, with unresponsive GRNs producing little to no

improvement. To test for GRNs’ response strength independently of my modeling effort,

I delivered 4 trials of a 1 second square pulse of tastant to every GRN before the

temporally rich stimulus. As above, I smoothed the trial-averaged spiking activity with a

Gaussian filter (s.d. = 60 ms.) to estimate the GRN’s PSTH. Also as above, ‘response

strength’ was defined as the absolute value of the difference between the mean firing rate

during the stimulus time and the background activity. The distribution of likelihood

improvements vs response strength was strongly bimodal (ΔBIC for Gaussian mixture

model fit with 2 components vs a 1 component model was >97). As shown in Fig. 15 one

mode centered on a cluster of largely unresponsive GRNs and showed little to no

improvement over the null model (mean = 1.07 Hz, -0.05 bits/spike improvement). A

second mode centered on responsive GRNs and showed a large improvement over the

null model (mean = 24.39 Hz, 0.74 bits/spike improvement).

69

Page 77: Gustatory Information Processing By Sam Reiter Thesis

Figure 15: The LLX improvement of a GRN model vs the response strength of the GRN. Each blue dot represents one GRN model. Marginal distribution histograms are plotted on either side of the scatterplot. The mean LLX improvement is plotted as the vertical dotted line. The component means of a 2 component Gaussian mixture model are plotted as green ‘Xs’. The GRN used to generate Fig.14 is plotted as the red dot.

As I used the same stimulus and type of nonlinearity for all the LNM fits, the

different relationships that GRNs have with tastant presentation can be compactly

depicted by the time course of LNM’s linear filter, k. GRNs were capable of responding

quickly and reliably to transient stimuli (filters peaked in 322.03 +/- 35.57 ms., n=69). A

single tastant was filtered differently by different GRNs and different tastants were

filtered differentially by single GRNs, together accounting for the previously observed

diversity of responses within the GRN population (Fig. 16).

70

Page 78: Gustatory Information Processing By Sam Reiter Thesis

Figure 16: GRNs exhibit a diversity of temporal receptive fields. (A) The jackknife mean +/- s.e. of 6 linear filters from models fit to 6 different GRNs all responding to the same stimulus (sucrose). The black filter is the one used to generate the data in Fig. 14. (B) GRN temporal filtering is tastant specific. The jackknife mean +/- s.e. of linear filters from models fit using a single GRN’s response to 2 different tastants. 200 ms. scale bar.

Thus, GRNs act as a bank of tastant specific rectified temporal filters of the time-

varying tastant concentration. This causes different tastants to activate different, but

overlapping, subsets of the GRN population with tastant specific patterns of spiking

activity.

A

B

71

Page 79: Gustatory Information Processing By Sam Reiter Thesis

2e. Estimating the information content of the GRN population

Representing tastants in this way should allow for information about tastant

identity and concentration to be transmitted to downstream neurons. To test this idea in a

quantitative way, I estimated the information about tastant identity contained within GRN

activity by using a simple nearest neighbor algorithm to discriminate between different

tastants given GRN population activity. The spike rate of each neuron in the analysis was

first estimated by smoothing the single trial spiking activity with a Gaussian filter (s.d. =

60 ms.). The mean of the first second of each trial (background activity) was then

removed. This was done to prevent any slow changes in spike rate over trials, or spiking

activity that “wrapped around” into the following trial, to be used to produce above

chance classification performance before tastant stimulation. This manipulation lowered

the baseline classification success to chance, while not affecting the peak discrimination

performance. (For example, for 12 way classification, 58.33 % accuracy after 500 ms.,

75.00% after 1.5s without background removal, compared to the 62.50% accuracy 500

milliseconds after tastant application, and 77.08 % after 1.5s reported in main text). The

dimensionality of the dataset was then reduced using Principal Component Analysis, with

the number of components chosen to preserve >90% of the dataset’s variance.

I performed K nearest neighbor classification, K=1. To generate confidence

intervals I used cross validation. I ran the classifier using variable lengths of data, to

observe how the ability to classify tastants evolves from the time of tastant presentation.

As a control test, I performed the same classification task using times when no tastant

was present.

72

Page 80: Gustatory Information Processing By Sam Reiter Thesis

In a 12 way discrimination task using the tastant set and data from Fig. 7, tastant

identity could be classified at 62.50% accuracy 500 milliseconds after tastant application

and 77.08 % accuracy 1.5 second after application (Fig. 16)

Figure 17: Tastant specific information is contained in the GRN population. (Blue) Cross validated success rate of 12 way K-nearest neighbor (K=1) classification of tastants using variable lengths of GRN population data (n=83), beginning at the time of tastant delivery. (Purple) As Blue, but beginning the data vector 2s before tastant delivery. Both show mean +/- 95% confidence interval. (Black line) Chance classification rate.

Classifying tastant identity using GRN population activity in this way represents a

lower bound on the amount of information present in the GRN population. It is probable

that with an increased number of GRNs, and with a more sophisticated classification

procedure, an observer could decode GRN activity with even higher accuracy. However,

the presence of tastant specific information within GRNs tells us nothing about whether

postsynaptic neurons make use it. Determining that requires studying second order

gustatory neurons directly.

73

Page 81: Gustatory Information Processing By Sam Reiter Thesis

2f. Second order gustatory neurons

Second order gustatory neurons (SONs) have not been identified in any insect. To

do so, I made simultaneous extracellular recordings of GRNs and sharp intracellular

recordings from putative SONs in the the sub-esophageal zone (SEZ). To determine

synaptically connected neurons, I calculated the spike triggered average (STA) between

pairs of GRNs and SONs. This was done by averaging the SON membrane potential

triggered on a GRN spike. The presence of SON spikes immediately following a GRN

spike would mask any excitatory postsynaptic potential (EPSP) waveform in the STA. I

therefore only considered GRN spikes that were not followed by an SON spike within

200 ms. Tastant delivery sometimes induced strong variability in the SON membrane

potential. To minimize the risk of this masking an EPSP waveform, I also did not include

any GRN spikes occurring during the square pulse tastant stimulation. Some GRNs were

completely silent unless activated by a tastant. To allow for a STA to be assessed for

these GRNs, I opened my tastant valve without applying pressure, providing a weak

stimulus sufficient to sporadically activate otherwise silent GRNs, but not strong enough

to induce strong SON membrane potential variability.

STAs were judged to indicate monosynaptic connections if a significant short

latency EPSP waveform was seen (Fig. 18). This was defined as a waveform that crossed

5 standard deviations from the background activity calculated from the average 100 ms.

prior to GRN spiking (8/70 pairs passed). The start of the EPSP waveform was estimated

as the time point where the STA crossed 3 standard deviations above the background.

This was required to occur within 5 milliseconds of the GRN spike (7/8 pairs passed).

The mean latency of these pairs was 2.80 +/- 0.28 ms., consistent with monosynaptic

74

Page 82: Gustatory Information Processing By Sam Reiter Thesis

synaptic delays176. Tastants used for paired recordings were sucrose (100 mM and 1000

mM), NaCl (100 mM and 1000 mM), caffeine (10 mM and 100 mM) and H2O.

Figure 18: A monosynaptic EPSP seen in a spike triggered average. EPSP waveform seen in the average membrane potential of a SON triggered on the spikes of a simultaneously recorded GRN (dashed line). Black trace is the mean, green traces are individual spike triggered membrane potential traces. Time is relative to the GRN spike.

Ideally, I would have filled the second order neurons identified through the above

STA analysis to examine the anatomy of SONs. The technical difficulty of that

experiment made it impossible. Instead, I performed many experiments recording from

neurons in the SEZ (without the simultaneous GRN recordings). Often, I could achieve

both physiological recording and anatomical reconstruction of the recorded neuron. For

intracellular dye filling, I used Neurobiotin (Vector labs), later conjugated to Alexa Fluor

488 or 568 (Molecular Probes). I imaged the neurons on a confocal microscope (Zeiss

510 inverted). Luckily for me, my blind sharp recordings from the ventral SEZ revealed

only 2 general classes of neurons. One type, descending neurons (DNs), extended an

75

Page 83: Gustatory Information Processing By Sam Reiter Thesis

axon down the Cervical Connective towards the body ganglia, and possessed stereotyped

sub-threshold membrane potential activity, spike shape, and baseline spike rate. After

n=16 recordings and fills, I was able to use these features to reliably identify DNs from

their physiology alone. The other type of neuron had physiology very similar to the

second order neurons identified by paired recordings (n=12 recordings and fills, see

below). Therefore all recorded neurons not identified as DNs were analyzed as SONs.

SONs had cell bodies in the SEZ, and most extended processes to bilaterally

cover the region where the GRNs project in the anterior SEZ (Fig. 19). Often their

processes continue beyond this region, into more the dorsal and posterior SEZ (Fig. 20).

Images of SONs were processed using FluoRender software (University of Utah). To

increase the clarity of the images, SONs were extracted from the original image z-stacks

through signal thresholding. These extracted images were then superimposed onto the

original Z stack, displayed at a dimmer color.

Figure 19: Anatomy of a SON. Frontal view of the moth brain, showing the anatomy of a single SON filling the anterior SEZ (z-stack). Compare with GRN projections shown in Fig. 5.

––

76

Page 84: Gustatory Information Processing By Sam Reiter Thesis

Figure 20: SON processes extend beyond the GRN projections. (A) Transverse view (down: anterior, up: posterior) of the SEZ. (Red) bilateral dextran fills of the GRNs of the Maxillary Nerves. (Green) Single SON, sending process lateral and posterior to GRN input. (B) Frontal view of the moth brain. A single filled SON is seen to send processes far anterior to the SEZ. Compare with GRN projections shown in Fig. 5.

My paired recordings revealed that SONs respond to different tastants than their

presynaptic GRNs. As in my analysis of GRNs, I examined a range of response

thresholds to demonstrate the robustness of my results. (for 5 s.d. threshold, GRNs

respond to 42.86% of tastants, SON to 77.55% of, correlation coefficient between

response profiles = 0.17, n=7). Even highly selective GRNs were found to synapse onto

broadly responsive SONs. For example, Fig. 21 shows a GRN that responds selectively to

sucrose. It synapses (STA EPSP shown in Fig. 17) onto a SON that responds to most

delivered tastants with tastant specific patterns of activity. This suggests that SONs

integrate the activity of multiple types of GRNs.

–– ––

A B

77

Page 85: Gustatory Information Processing By Sam Reiter Thesis

Figure 21: A narrowly tuned GRN synapses onto a broadly tuned SON. Raster plots of synaptically connected, simultaneously recorded GRN and SON responding to multiple trials of different tastants and concentrations. (Bottom left, Red) Average color sensor voltage trace. Tastant delivery time is also marked with shading. (Bottom right, Blue) Voltage trace of the SON during the last trial of the raster plot. The STA of Fig. 18 was generated from these two neurons.

As a population, SONs responded to more tastants than GRNs (Fig. 22 a-c, c.f.

Fig. 22 d-f, Chi squared tests between GRNs and SONs at every response threshold,

p<0.05 after Bonferroni correction). While different tastants activated different numbers

of SONs (Fig. 22b), sucrose did not consistently activate significantly more SONs than

any other delivered tastants (at an 8 s.d. threshold, Chi squared tests, p>0.05 after

Bonfferoni correction), and very few SONs were selective for any delivered tastant. (Fig.

22c, mean across all thresholds = 0.19+/-1.2 %).

78

Page 86: Gustatory Information Processing By Sam Reiter Thesis

Figure 22: (A-C) Summary statistics of SON population data. (A) The percent of the SON population that responded to varying numbers of tastants. (n=13) (B) The percent of the SON population that responded to the listed tastant. (C) The percent of neurons that responded to the listed tastant and do not respond to any other delivered tastant. (D-F) Summary statistics of GRN population data responding to the same tastant set as delivered to SONs (n=83). Plots as A-C.

Given the temporal structure of their GRN input, SON tastant responses could be

quite temporally complex, with responses to square stimuli often reliably eliciting

sequences of excitation and inhibition that continue past the time of tastant delivery (Fig.

23). To build intuition about the meaning of the various response thresholds of Fig. 22, I

numbered the maximum of threshold (1-12) in which each block of trials qualifies as a

79

Page 87: Gustatory Information Processing By Sam Reiter Thesis

response. For example, GRN 1’s strong response to lobeline counts as a response for

thresholds 1-12, but its weaker response to sucrose only counts as a response up to

threshold 5. Because it is not known how neurons postsynaptic to these filter their SON

input, choosing what constitutes a response is subjective. However, in the absence of

evidence to the contrary a rational approach is to use a lower threshold, as it doesn’t

ignore potentially informative variability in SON spiking.

Figure 23: The tastant responses of SONs. Raster plots of 4 example SONs responding to various tastants. Written to the right of every block of trials is the maximum threshold (1-12) in which that block of activity constitutes a response. (Bottom, Red) Average color sensor voltage trace. Tastant delivery time is also marked with shading. (Bottom, Blue) Voltage trace of the SON during the last trial of the raster plot. Scale bars are 10 mV.

While the filtering properties of GRNs resulted in the average GRN response

mimicking the temporal profile of the stimulus, the average SON response quickly

peaked at the onset and, to a lesser extent, the offset of the stimulus (Fig. 24).

GRN 1 GRN 2 GRN 3 GRN 4

80

Page 88: Gustatory Information Processing By Sam Reiter Thesis

Figure 24: SONs transform their input temporally. The average spiking response of GRNs (Blue, n=83) and SONs (Black, n=13) in response to 1 second application of 7 different tastants (those of Fig. 22). The average color sensor voltage trace is shown in red. Spiking responses were smoothed with a Gaussian filter (s.d. 60 ms.).

Like GRNs, SON responses depended on, and therefore contained information

about, tastant identity. Using the temporal responses across a population of second order

cells, my nearest neighbor algorithm could identify tastant identity at 96.43 +/- 3.57%

accuracy 500 milliseconds and 100 +/- 0% 1.5 second after tastant application (7 way

discrimination, Fig. 25).

81

Page 89: Gustatory Information Processing By Sam Reiter Thesis

Figure 25: Tastant specific information is contained in the SON population. (Blue) Cross validated success rate of 12 way K-nearest neighbor (K=1) classification of tastants using variable lengths of SON population data (n=13), beginning at the time of tastant delivery. (Purple) As Blue, but beginning the data vector 2s before tastant delivery. Both show mean +/- 95% confidence interval. (Black line) Chance classification rate.

A consequence of the spatial and temporal transformations of tastant information

as it is passed from GRNs to SONs is that tastants could be discriminated at a higher

accuracy more quickly, using fewer neurons, in SONs than in GRNs (Fig. 26).

Information about tastant identity cannot be present in the SON population if it isn’t

already present in the GRNs. This suggests that because SONs integrate the activity of

multiple types of GRNs, recording the activity of a small number of SONs becomes

equivalent to recording the activity of a very large set of GRNs.

Figure 26: SONs classify tastants more efficiently than GRNs. Comparison of 7 way classification using the GRN population (n=83, green) and the SON population (n=13, red). Both show mean +/- 95% confidence interval. (Black line) Chance classification rate.

82

Page 90: Gustatory Information Processing By Sam Reiter Thesis

3. DISCUSSION

Using a simple model system, I studied in detail how tastants are represented in

the population of GRNs, and how this information is processed by a postsynaptic neural

population. As we have seen, gustatory physiology has often been interpreted in terms of

‘basic tastes’ (BTs). In this paradigm, chemicals are associated with one or another BT

category, and the strategy for representing these categories is studied.

I was unable to interpret my results in terms of the BTs. Many GRNs responded

to some chemicals in a category, but not to others. Other GRNs responded to chemicals

associated with multiple categories. GRNs synapsed onto SONs that responded to tastants

in different categories than their presynaptic GRNs, and themselves were broadly

responsive across categories. In addition, moths produced different behaviors when

stimulated by chemicals associated with the same BT.

Rather than representing BT categories, either through dedicated neurons

(‘labeled lines’) or across a population of neurons (‘across fiber patterns’), I found that

the GRN and SON population contained tastant specific information. Different tastants

caused different but overlapping sets of neurons to respond with tastant specific temporal

patterns of activity. The tastant specific behavior of the moth suggests that neurons

downstream from GRNs and/or SONs make use of this information.

I propose that instead of representing a small number of tastant categories, the

gustatory system functions to uniquely represent a large range of chemicals. Although

often interpreted in terms of BTs, tastant specific human perceptions93,156, animal

behaviors81,131,157, and cellular responses131,133,149 have been reported in many animals. I

83

Page 91: Gustatory Information Processing By Sam Reiter Thesis

believe that removing the BT interpretation from these and other studies clarifies the

findings, often without calling experimental results into question. The complexity of

gustation at all these levels makes it likely that any attempt to replace the BTs with some

other simple category scheme will also fail. More generally, a combinatorial code is

widely accepted as being used to uniquely represent a large range of stimuli in every

sensory system besides gustation5,6. My results suggest that gustation is no exception.

Do basic tastes exist in any context? As with all eliminative theories177, one can

take mine too far. I am not advocating that we remove basic taste words from everyday

use. The fact that we have inherited BT terms from thousands of years of humans

describing their gustatory perceptions suggests that the BTs align with human perception

much more than some arbitrary partition of tastant chemicals into 4 or 5 categories. What

I am advocating is that we do not rely on the BTs to define the limits of gustatory

perception. I see the BTs as rough umbrella terms that each cover areas of a high

dimensional perceptual space. As there are many tastants that cannot be described in

terms of BTs, the BTs do not constitute a basis set. And because some chemicals are

often described as having a taste associated with multiple BTs, we should view the BTs

as applying to partially overlapping subsets of perceptual space. In short I think about

using BT terms as similar to everyday words used to describe odors. We know that when

we describe a smell as “fruity”, “earthy”, or “savory”, we are not defining olfactory

perception exactly.

I am much more critical of applying the BTs to non-human animals. Here it seems

that there is little to be gained from the simplicity of the BT terms, and much to be lost.

The BT model obscures the wonderful diversity of animal gustatory behaviors. For

84

Page 92: Gustatory Information Processing By Sam Reiter Thesis

example, my results show that moths extend their proboscis (generally considered an

attractive behavior) in response to sucrose, but not to other sugars. I think of the task of

gustatory neuroscience as one of explaining this diversity, and so employing BT

terminology to describe the neural basis of behavior at best oversimplifies, and at the

worst misleads.

Removing the BT interpretation from gustatory research and coming to terms

combinatorial coding is just one step towards understanding the gustatory system. The

next step is to determine the specific computations that the gustatory system performs. I

consider the modeling approach I applied to studying GRNs as a useful step along this

path. Hopefully future work will attempt to build encoding models of higher order

gustatory neurons. I view my study of SONs as a sketch of such a model, with the tastant

responses of SONs being viewed in terms of their GRN input. A proper encoding model

of SONs would take the form of a cascade of LNMs, with the tastant, passed through the

bank of GRN filters, acting as the input to an SON model. Such cascade models are an

area of active research168,178.

Of equal importance to building an accurate model of how tastant representations

are transformed as they move from one neural population to another is understanding the

reason such transformations take place. While the function of GRNs can be thought of as

transducing the presence of tastants into tastant specific information, SONs remain

largely mysterious. I have described how SON’s become more broadly responsive to

tastants by integrating the activity of multiple types of GRNs. SONs also respond with

seemingly more elaborate temporal patterns of spikes. This serves to make single SONs

more informative about tastant identity at faster speeds than single GRNs (Fig. 25). This

85

Page 93: Gustatory Information Processing By Sam Reiter Thesis

is notable because it directly opposes the intuition that more narrowly tuned neurons are

able to provide an unambiguous signal faster than more broadly tuned neurons. It is

probable that these transformations possess additional significance that is presently

unclear. A fuller understanding of SONs will require the identification of 3rd order

gustatory neurons; only by studying how activity in a neural population is used by

postsynaptic population(s) can we accurately infer its function(s). Understanding 3rd

order gustatory neurons will in turn require an understanding of fourth order gustatory

neurons, which requires an understanding of fifth order neurons….Until we can build and

understand an encoding model of behavior, understood in terms of motor neurons,

understood in terms of interneurons, understood in terms of GRNs, we cannot say that we

truly understand gustatory information processing. There is a lot of work to be done.

One difficulty of coming to terms with combinatorial coding in the gustatory

system is that now, like in olfaction, we must admit that we do not know the

dimensionality of tastant space. A true understanding of either system will require us to

be able to predict how some new, never before tested, chemical will be represented in the

brain. This is almost trivial in low level vision or audition, where we know the stimulus

dimensionality. For example, if asked to predict the visual system’s response to some

novel stimulus, I might ask for the stimulus’ wavelength, intensity and position in retinal

coordinates. Knowing roughly how photoreceptors at different points in the retina

respond to such things, I would have some prediction of the initial visual response.

Presently, knowing how GRNs respond to sucrose, maltose, NaCl and caffeine doesn’t

tell me how they will respond to some new chemical, or mixture of chemicals.

86

Page 94: Gustatory Information Processing By Sam Reiter Thesis

While some efforts of defining perceptual space have been made in both the

olfactory179 and gustatory26 system, the vast number of chemicals, each of which can be

perceived as different when used at different concentrations, or when combined in

mixtures, has prevented the creation of the kind of comprehensive model needed to build

encoding models in terms of perceptual space. And even if such a model were

constructed, it could not be applied to other animals. Thus, providing a systematic

account of these spaces seems very difficult.

Nevertheless, it seems likely that the perceptual space of tastants is much lower

than the number of stimuli that the system can represent. Indeed, it would be fascinating

to find sets of tastants that, presented at appropriate concentrations, always elicit the same

behaviors. For a given behavior, these sets will certainly be species specific. But if such

sets could be found, then it may be useful to describe the strategies used by the animal’s

nervous system to represent such tastants. In some ways this project would be similar to

an analysis of how the BTs are represented by the brain. But it would differ in several key

respects. It would acknowledge the species specificity of any findings. It would define

the mapping between tastants and behavioral categories after observing many behaviors,

not after observing a single behavior or establishing the mapping a priori. And it would

put these findings in terms of regularities of the combinatorial code for tastants.

Therefore, this project represents a logical next step, only possible after first establishing

the general code used by the gustatory system.

Lastly, the luckiest move I made in my work was to choose to study gustation in

moths. It is hard to think of a better animal in which to study the first stages of the

gustatory system: Manduca possess an extremely long proboscis, hold up very well

87

Page 95: Gustatory Information Processing By Sam Reiter Thesis

during long experiments, do not need to be anesthetized, are large enough to make

routine extracellular recordings from a nerve, the list goes on and on. It seems to me that

the price of not having as many general tools, like the ability to do genetic manipulations,

in this case has been more than overcome by the advantages of choosing this animal for

my specific problem. I imagine this could be a much more general phenomenon, where

every biological question is most easily answered by studying a particular animal model.

I doubt that this animal will always be a fly, mouse, fish, or monkey.

Even if this is not the case, and it is uniformly more difficult to study atypical model

animals, I think it is still worthwhile. For how can we tell a general principle of neural

information processing from a species specific quirk without studying and comparing a

wide range of animals?

88

Page 96: Gustatory Information Processing By Sam Reiter Thesis

4. APPENDIX: SPIKE SORTING

Data was collected at 40 kHz (LabView software), amplified 3000 times, and

band pass filtered 300 Hz to 6000 Hz. Following the method of Pouzat et al.160, spike

sorting was performed by first building a set of template models representing every

neuron that was extracellularly recorded during an experiment, and then using these

models to classify all potential spiking events found in the recording. In order to build

the template models, the recorded data was preprocessed by smoothing with a 3 point box

filter, and normalized by channel specific mean absolute deviation. Spikes were initially

detected by extracting times where the preprocessed data crossed a user specified,

channel specific amplitude threshold. Putative spike waveforms were then extracted

using a user-defined temporal window to either side of the threshold crossing.

Waveforms were concatenated across channels, creating a single dimensional spike

waveform vector of length (number of samples) by (number of channels). All event

waveforms were individually “dejittered” around the mean event waveform event at sub-

sampling resolution so that deviations from the mean were not simply a result of our

sampling frequency180.

In order to ensure that all waveforms in the template models corresponded to the

electrical activity of single neurons (and not, for example, the superposition of

waveforms from two different neurons), we excluded a small number of putative

superposition waveforms prior to further analysis. Waveforms generated from the

superposition of multiple neurons were expected to have peaks on either side of the

principle spike peak. To identify and exclude such superpositions, we defined a number

of standard deviations away from the median spike waveform. All waveforms that exceed

89

Page 97: Gustatory Information Processing By Sam Reiter Thesis

this threshold were excluded as potential superpositions. For the remaining 'clean' data,

the spike waveform dimensionality was reduced using principle component analysis

(PCA) to allow for efficient clustering. We selected the number of principal components

(generally more than three) to project the data onto, using the shoulder of the scree plot

(cumulative proportion of variance accounted for as a function of the number of

eigenvectors used) as a guide. After looking at the event data projected onto the three-

dimensional subspace described by the three largest principle components, a user-defined

range of clusters was selected to consider for model selection.

Clustering of the event data was accomplished by using the expectation

maximization algorithm, for the user-defined range of clusters. The Bayesian Information

Criterion181 was used to perform model selection. Treating this automated clustering as a

starting point, we merged, split, or removed clusters as necessary, using several visual

projections of the data to assess the result of these operations. Three methods of visually

evaluating putative clusters included projecting the clustered data onto the first 3

principle components, visually inspecting the waveforms for each cluster, and projecting

each pair of clusters onto the 2 principle components that maximally separate that pair.

After reaching a satisfying set of cluster labels, the top 20% of outliers based on the

Mahalonobis distance for each cluster were removed in order to ensure that cluster means

were not biased by outlier waveforms.

The resultant mean waveforms (model centers) and an estimate of the noise

distribution of each recording channel were used to classify all spikes in the dataset, by

selecting the cluster mean (or superposition of two cluster means, see below) which

minimized the Mahalanobis distance of the residual waveform from the channel noise.

90

Page 98: Gustatory Information Processing By Sam Reiter Thesis

Assuming Gaussian noise, the distribution of residuals (putative spiking event-cluster

mean) across all clusters was expected to follow a chi squared distribution. Spike

waveforms corresponding to residuals above the estimated chi-square 1% value were

marked as potential superpositions. Such waveforms were compared to a supermodel

containing every pair of model centers shifted at every possible time lag relative to each

other, as well as the original model centers. The potential superposition waveforms were

matched with their closest supermodel entries by again determining the model center that

left a minimal distance between the resulting residual and the estimation of the channel

noise.

After removing outliers , defined as the top 1% of from each cluster using

Mahalonobis distance, the classified clusters were evaluated visually for quality, with

consideration given to the standard deviation of the channel noise following removal of

putative spiking events, and the pairwise separation of clusters as determined by the

distance between means when projected along the axis defined by the Fischer linear

discriminant (FLD)160. Cluster quality was also quantitatively evaluated by estimating the

false positive and false negative rates162, composed of: the percent of spikes in each

cluster with waveforms falling below the threshold limit, the percentage of spikes

misclassified as belonging to another cluster, the percentage of spikes incorrectly

assigned to the current cluster, and the percentage of spikes that fall within an un-

physiological inter spike interval (e.g. < 2 ms.). We typically required that the associated

type 1 and type 2 errors be below 5% in order for a cluster to be considered well isolated.

91

Page 99: Gustatory Information Processing By Sam Reiter Thesis

4. BIBLIOGRAPHY

1. Young, T. The Bakerian Lecture: On the Theory of Light and Colours. Philos.

Trans. R. Soc. Lond. 92, 12–48 (1802).

2. Young, T. A course of lectures on natural philosophy and the mechanical arts.

(1807).

3. Erickson, R. A study of the science of taste: On the origins and influence of the core

ideas. Behav. Brain Sci. (2008).

4. Victor, J. D. & Nirenberg, S. Indices for Testing Neural Codes. Neural Comput. 20,

2895–2936 (2008).

5. Mcllwain, J. Population coding: a historical sketch. Prog. Brain Res. 130, 3–7

(2001).

6. Erickson, R. The evolution and implications of population and modular neural

coding ideas. Prog. Brain Res. 130, 9–29 (2001).

7. Müller, J. Handbuch der Physiologie des Menschen für Vorlesungen. (Verlag von J.

Hölscher, 1837).

8. Boring, E. Sensation and perception in the history of experimental psychology. (D.

Appleton-Century Company, 2014).

9. Rose, D. The historical roots of the theories of local signs and labelled lines.

Perception 28, 675–685 (1999).

10. Yarmolinsky, D. A., Zuker, C. S. & Ryba, N. J. P. Common Sense about Taste:

From Mammals to Insects. Cell 139, 234–244 (2009).

92

Page 100: Gustatory Information Processing By Sam Reiter Thesis

11. Bushdid, C., Magnasco, M. O., Vosshall, L. B. & Keller, A. Humans Can

Discriminate More than 1 Trillion Olfactory Stimuli. Science 343, 1370–1372

(2014).

12. Gilad, Y., Man, O. & Glusman, G. A comparison of the human and chimpanzee

olfactory receptor gene repertoires. Genome Res. 15, 224–230 (2005).

13. Stensmyr, M. C. et al. A Conserved Dedicated Olfactory Circuit for Detecting

Harmful Microbes in Drosophila. Cell 151, 1345–1357 (2012).

14. Kurtovic, A. et al. A single class of olfactory neurons mediates behavioural

responses to a Drosophila sex pheromone. Nature 446, 542–546 (2007).

15. Unschuld, P. Huang Di Nei Jing Su Wen: Nature, Knowledge, Imagery in an

Ancient Chinese Medical Text. (University of California Press, 2003).

16. Aristotle. On the Soul. (1931).

17. Erickson, R. P. Ohrwall, Henning and von Skramlik; the foundations of the four

primary positions in taste. Neurosci. Biobehav. Rev. 8, 105–27 (1984).

18. Delwiche, J. Are there ‘basic’tastes? Trends Food Sci. Technol. 7, 411–415 (1996).

19. Berlin, B. & Kay, P. Basic Color Terms: Their Universality and Evolution.

(University of California Press, 1969).

20. Jameson, K. Culture and cognition: What is universal about the representation of

color experience? J. Cogn. Cult. 3-4, 293–347 (2005).

21. Schiffman, S. & Erickson, R. The issue of primary tastes versus a taste continuum.

Neurosci. Biobehav. Rev. (1980).

22. Erickson, R. P. Studies on the perception of taste: Do primaries exist? Physiol.

Behav. 28, 57–62 (1982).

93

Page 101: Gustatory Information Processing By Sam Reiter Thesis

23. Mcburney, D., Smith, D. & Shick, T. Gustatory cross adaptation: sourness and

bitterness. Percept. Psychophys. (1972).

24. Breslin, P. A. S. & Beauchamp, G. K. Suppression of Bitterness by Sodium:

Variation Among Bitter Taste Stimuli. Chem. Senses 20, 609–623 (1995).

25. Spector, A. C. & Kopka, S. L. Rats fail to discriminate quinine from denatonium:

implications for the neural coding of bitter-tasting compounds. J. Neurosci. Off. J.

Soc. Neurosci. 22, 1937–1941 (2002).

26. Schiffman, S., Mcelroy, A. & Erickson, R. The range of taste quality of sodium

salts. Physiol. Behav. (1980).

27. Bartoshuk, L. M., Murphy, C. & Cleveland, C. T. Sweet taste of dilute NaCl:

Psychophysical evidence for a sweet stimulus. Physiol. Behav. 21, 609–613 (1978).

28. Schiffman, S. S. et al. The effect of sweeteners on bitter taste in young and elderly

subjects. Brain Res. Bull. 35, 189–204 (1994).

29. Schiffman, S. S., Sennewald, K. & Gagnon, J. Comparison of taste qualities and

thresholds of D- and L-amino acids. Physiol. Behav. 27, 51–59 (1981).

30. Auvray, M. & Spence, C. The multisensory perception of flavor. Conscious. Cogn.

17, 1016–1031 (2008).

31. Pfaffmann, C. Gustatory afferent impulses. J. Cell. Comp. Physiol. 17, 243–258

(1941).

32. Feigin, M., Sclafani, A. & Sunday, S. Species differences in polysaccharide and

sugar taste preferences. Neurosci. Biobehav. (1987).

33. Carpenter, J. A. Species differences in taste preferences. J. Comp. Physiol. Psychol.

49, 139–144 (1956).

94

Page 102: Gustatory Information Processing By Sam Reiter Thesis

34. Baker, H. & Baker, I. A brief historical review of the chemistry of floral nectar.

(Columbia University Press, 1983).

35. Glendinning, J. Preference and aversion for deterrent chemicals in two species of

Peromyscus mouse. Physiol. Behav. (1993).

36. Glendinning, J. I. Is the bitter rejection response always adaptive? Physiol. Behav.

56, 1217–1227 (1994).

37. Glendinning, J. How do predators cope with chemically defended foods? Biol. Bull.

(2007).

38. Pfaffmann, C. Taste and Smell. Annu. Rev. Psychol. 7, 391–408 (1956).

39. King, M. The Role of the Nucleus of the Solitary Tract in Gustatory Processing.

Chapter 2 Anatomy of the Rostral Nucleus of the Soltiary Tract. (CRC Press).

40. Whitehead, M. & Frank, M. Anatomy of the gustatory system in the hamster:

Central projections of the chorda tympani and the lingual neirve. J. Comp. Neurol.

(1983).

41. Lundy, R. & Norgren, R. Pontine gustatory activity is altered by electrical

stimulation in the central nucleus of the amygdala. J. Neurophysiol. (2001).

42. Reilly, S. The role of the gustatory thalamus in taste-guided behavior. Neurosci.

Biobehav. Rev. 22, 883–901 (1998).

43. Yamamoto, T. & Yuyama, N. On a neural mechanism for cortical processing of

taste quality in the rat. Brain Res. 400, 312–320 (1987).

44. Yamamoto, T., Yuyama, N., Kato, T. & Kawamura, Y. Gustatory responses of

cortical neurons in rats. I. Response characteristics. J. Neurophysiol. 51, 616–635

(1984).

95

Page 103: Gustatory Information Processing By Sam Reiter Thesis

45. Nakamura, K. & Norgren, R. Gustatory responses of neurons in the nucleus of the

solitary tract of behaving rats. J. Neurophysiol. 66, 1232–1248 (1991).

46. Nishijo, H. & Norgren, R. Parabrachial gustatory neural activity during licking by

rats. J. Neurophysiol. 66, 974–985 (1991).

47. Frank, M. An analysis of hamster afferent taste nerve response functions. J. Gen.

Physiol. 61, (1973).

48. Pritchard, T. C., Hamilton, R. B. & Norgren, R. Neural coding of gustatory

information in the thalamus of Macaca mulatta. J. Neurophysiol. 61, 1–14 (1989).

49. Rolls, E. T., Yaxley, S. & Sienkiewicz, Z. J. Gustatory responses of single neurons

in the caudolateral orbitofrontal cortex of the macaque monkey. J. Neurophysiol. 64,

1055–1066 (1990).

50. Scott, T. R., Plata-Salaman, C. R., Smith, V. L. & Giza, B. K. Gustatory neural

coding in the monkey cortex: stimulus intensity. J. Neurophysiol. 65, 76–86 (1991).

51. Smith, D. V. & Hanamori, T. Organization of gustatory sensitivities in hamster

superior laryngeal nerve fibers. J. Neurophysiol. 65, 1098–1114 (1991).

52. Di Lorenzo, P. M. & Monroe, S. Taste responses in the parabrachial pons of male,

female and pregnant rats. Brain Res. Bull. 23, 219–227 (1989).

53. Lemon, C. H. & Katz, D. B. The neural processing of taste. BMC Neurosci. 8, S5

(2007).

54. Katz, D. B., Simon, S. A. & Nicolelis, M. A. L. Dynamic and Multimodal

Responses of Gustatory Cortical Neurons in Awake Rats. J. Neurosci. 21, 4478–

4489 (2001).

96

Page 104: Gustatory Information Processing By Sam Reiter Thesis

55. Lemon, C. H. & Smith, D. V. Influence of Response Variability on the Coding

Performance of Central Gustatory Neurons. J. Neurosci. 26, 7433–7443 (2006).

56. Di Lorenzo, P. M. & Victor, J. D. Taste Response Variability and Temporal Coding

in the Nucleus of the Solitary Tract of the Rat. J. Neurophysiol. 90, 1418–1431

(2003).

57. Chen, J.-Y., Victor, J. D. & Di Lorenzo, P. M. Temporal coding of intensity of NaCl

and HCl in the nucleus of the solitary tract of the rat. J. Neurophysiol. 105, 697–711

(2011).

58. Rosen, A. M., Victor, J. D. & Di Lorenzo, P. M. Temporal coding of taste in the

parabrachial nucleus of the pons of the rat. J. Neurophysiol. 105, 1889–1896 (2011).

59. Roussin, A. T., D’Agostino, A. E., Fooden, A. M., Victor, J. D. & Di Lorenzo, P.

M. Taste coding in the nucleus of the solitary tract of the awake, freely licking rat. J.

Neurosci. Off. J. Soc. Neurosci. 32, 10494–10506 (2012).

60. Weiss, M. S., Victor, J. D. & Di Lorenzo, P. M. Taste coding in the parabrachial

nucleus of the pons in awake, freely licking rats, and comparison with the nucleus of

the solitary tract. J. Neurophysiol. (2013).

61. Halpern, B. P. & Tapper, D. N. Taste Stimuli: Quality Coding Time. Science 171,

1256–1258 (1971).

62. Scott, T. R. Behavioral support for a neural taste theory. Physiol. Behav. 12, 413–

417 (1974).

63. Dahl, M., Erickson, R. . & Simon, S. . Neural responses to bitter compounds in rats.

Brain Res. 756, 22–34 (1997).

97

Page 105: Gustatory Information Processing By Sam Reiter Thesis

64. Danilova, V., Hellekant, G., Tinti, J.-M. & Nofre, C. Gustatory Responses of the

Hamster Mesocricetus auratus to Various Compounds Considered Sweet by

Humans. J. Neurophysiol. 80, 2102–2112 (1998).

65. Hellekant, G., Danilova, V. & Ninomiya, Y. Primate Sense of Taste: Behavioral and

Single Chorda Tympani and Glossopharyngeal Nerve Fiber Recordings in the

Rhesus Monkey, Macaca mulatta. J. Neurophysiol. 77, 978–993 (1997).

66. Lemon, C. & Smith, D. Neural representation of bitter taste in the nucleus of the

solitary tract. J. Neurophysiol. (2005).

67. Verhagen, J. V., Giza, B. K. & Scott, T. R. Responses to Taste Stimulation in the

Ventroposteromedial Nucleus of the Thalamus in Rats. J. Neurophysiol. 89, 265–

275 (2003).

68. Giza, B. K., Scott, T. R., Sclafani, A. & Antonucci, R. F. Polysaccharides as taste

stimuli: their effect in the nucleus tractus solitarius of the rat. Brain Res. 555, 1–9

(1991).

69. Michel, W. & Caprio, J. Responses of single facial taste fibers in the sea catfish,

Arius felis, to amino acids. J. Neurophysiol. 66, 247–260 (1991).

70. Scott, T. R. & Chang, F.-C. T. The state of gustatory neural coding. Chem. Senses 8,

297–314 (1984).

71. Smith-Swintosky, V. L., Plata-Salaman, C. R. & Scott, T. R. Gustatory neural

coding in the monkey cortex: stimulus quality. J. Neurophysiol. 66, 1156–1165

(1991).

72. Frank, M. E., Bieber, S. L. & Smith, D. V. The organization of taste sensibilities in

hamster chorda tympani nerve fibers. J. Gen. Physiol. 91, 861–896 (1988).

98

Page 106: Gustatory Information Processing By Sam Reiter Thesis

73. Frank, M. E. Taste-responsive neurons of the glossopharyngeal nerve of the rat. J.

Neurophysiol. 65, 1452–1463 (1991).

74. Scott, T. R. & Giza, B. K. Coding channels in the taste system of the rat. Science

249, 1585–1587 (1990).

75. Hastie, T., Tibshirani, R. & Friedman, J. The Elements of Statistical Learning: Data

Mining, Inference, and Prediction. (Springer, 2009).

76. Erickson, R. P., Rodgers, J. L. & Sarle, W. S. Statistical analysis of neural

organization. J. Neurophysiol. 70, 2289–2300 (1993).

77. Erickson, R. P., Doetsch, G. S. & Marshall, D. A. The Gustatory Neural Response

Function. J. Gen. Physiol. 49, 247–263 (1965).

78. Plata-Salaman, C. R., Scott, T. R. & Smith-Swintosky, V. L. Gustatory neural

coding in the monkey cortex: L-amino acids. J. Neurophysiol. 67, 1552–1561

(1992).

79. Erickson, R., Covey, E. & Doetsch, G. Neuron and stimulus typologies in the rat

gustatory system. Brain Res. (1980).

80. Woolston, D. & Erickson, R. Concept of neuron types in gustation in the rat. J.

Neurophysiol. (1979).

81. Glendinning, J. I. Temporal Coding Mediates Discrimination of ‘Bitter’ Taste

Stimuli by an Insect. J. Neurosci. 26, 8900–8908 (2006).

82. Wilson, D. M., Boughter, J. D. & Lemon, C. H. Bitter Taste Stimuli Induce

Differential Neural Codes in Mouse Brain. PLoS ONE 7, e41597 (2012).

83. Pfaffmann, C., Frank, M. & Norgren, R. Neural Mechanisms and Behavioral

Aspects of Taste. Annu. Rev. Psychol. 30, 283–325 (1979).

99

Page 107: Gustatory Information Processing By Sam Reiter Thesis

84. Frank, M. E. Neuron types, receptors, behavior, and taste quality. Physiol. Behav.

69, 53–62 (2000).

85. Hellekant, G., Ninomiya, Y. & Danilova, V. Taste in chimpanzees. III: Labeled-line

coding in sweet taste. Physiol. Behav. 65, 191–200 (1998).

86. Hettinger, T. P. & Frank, M. E. Information processing in mammalian gustatory

systems. Curr. Opin. Neurobiol. 2, 469–478 (1992).

87. Chen, X., Gabitto, M., Peng, Y., Ryba, N. J. P. & Zuker, C. S. A Gustotopic Map of

Taste Qualities in the Mammalian Brain. Science 333, 1262–1266 (2011).

88. Chaudhari, N. & Roper, S. D. The cell biology of taste. J. Cell Biol. 190, 285–296

(2010).

89. Roper, S. Regenerative impulses in taste cells. Science 220, 1311–1312 (1983).

90. Pérez, C. A. et al. A transient receptor potential channel expressed in taste receptor

cells. Nat. Neurosci. 5, 1169–1176 (2002).

91. Chandrashekar, J. et al. The cells and peripheral representation of sodium taste in

mice. Nature 464, 297–301 (2010).

92. Heck, G., Mierson, S. & DeSimone, J. Salt taste transduction occurs through an

amiloride-sensitive sodium transport pathway. Science 223, 403–405 (1984).

93. Hettinger, T. P., Gent, J. F., Marks, L. E. & Frank, M. E. A confusion matrix for the

study of taste perception. Percept. Psychophys. 61, 1510–1521 (1999).

94. Huang, A. L. et al. The cells and logic for mammalian sour taste detection. Nature

442, 934–938 (2006).

95. Vandenbeuch, A. & Kinnamon, S. C. Why do taste cells generate action potentials?

J. Biol. 8, 42 (2009).

100

Page 108: Gustatory Information Processing By Sam Reiter Thesis

96. Huang, Y.-J. et al. The role of pannexin 1 hemichannels in ATP release and cell-cell

communication in mouse taste buds. Proc. Natl. Acad. Sci. 104, 6436–6441 (2007).

97. Taruno, A. et al. CALHM1 ion channel mediates purinergic neurotransmission of

sweet, bitter and umami tastes. Nature 495, 223–226 (2013).

98. Finger, T. E. et al. ATP Signaling Is Crucial for Communication from Taste Buds to

Gustatory Nerves. Science 310, 1495–1499 (2005).

99. Kinnamon, J., Sherman, T. & Roper, S. Ultrastructure of mouse vallate taste buds:

III. Patterns of synaptic connectivity. J. Comp. Neurol. (1988).

100. Huang, Y. A., Dando, R. & Roper, S. D. Autocrine and Paracrine Roles for ATP

and Serotonin in Mouse Taste Buds. J. Neurosci. 29, 13909–13918 (2009).

101. Huang, Y.-J. et al. Mouse Taste Buds Use Serotonin as a Neurotransmitter. J.

Neurosci. 25, 843–847 (2005).

102. Cao, Y., Zhao, F. -l., Kolli, T., Hivley, R. & Herness, S. GABA expression in the

mammalian taste bud functions as a route of inhibitory cell-to-cell communication.

Proc. Natl. Acad. Sci. 106, 4006–4011 (2009).

103. Huang, Y. A., Maruyama, Y. & Roper, S. D. Norepinephrine Is Coreleased with

Serotonin in Mouse Taste Buds. J. Neurosci. 28, 13088–13093 (2008).

104. Ogura, T. et al. Immuno-localization of vesicular acetylcholine transporter in mouse

taste cells and adjacent nerve fibers: indication of acetylcholine release. Cell Tissue

Res. 330, 17–28 (2007).

105. Vandenbeuch, A. et al. Evidence for a role of glutamate as an efferent transmitter in

taste buds. BMC Neurosci. 11, 77 (2010).

101

Page 109: Gustatory Information Processing By Sam Reiter Thesis

106. Zaidi, F. N. Discrete Innervation of Murine Taste Buds by Peripheral Taste

Neurons. J. Neurosci. 26, 8243–8253 (2006).

107. Bear, M. F., Connors, B. W. & Paradiso, M. A. Neuroscience: Exploring the Brain.

(Lippincott Williams & Wilkins, 2007).

108. Adler, E. et al. A Novel Family of Mammalian Taste Receptors. Cell 100, 693–702

(2000).

109. Meyerhof, W. et al. The Molecular Receptive Ranges of Human TAS2R Bitter

Taste Receptors. Chem. Senses bjp092 (2009).

110. Chandrashekar, J. et al. T2Rs Function as Bitter Taste Receptors. Cell 100, 703–711

(2000).

111. Mueller, K. L. et al. The receptors and coding logic for bitter taste. Nature 434,

225–229 (2005).

112. Caicedo, A. & Roper, S. D. Taste receptor cells that discriminate between bitter

stimuli. Science 291, 1557–1560 (2001).

113. Nelson, G. et al. Mammalian Sweet Taste Receptors. Cell 106, 381–390 (2001).

114. Kitagawa, M., Kusakabe, Y., Miura, H., Ninomiya, Y. & Hino, A. Molecular

genetic identification of a candidate receptor gene for sweet taste. Biochem.

Biophys. Res. Commun. 283, 236–242 (2001).

115. Kim, M.-R. et al. Regional expression patterns of taste receptors and gustducin in

the mouse tongue. Biochem. Biophys. Res. Commun. 312, 500–506 (2003).

116. Nelson, G. et al. An amino-acid taste receptor. Nature 416, 199–202 (2002).

117. Zhao, G. Q. et al. The Receptors for Mammalian Sweet and Umami Taste. Cell 115,

255–266 (2003).

102

Page 110: Gustatory Information Processing By Sam Reiter Thesis

118. Chandrashekar, J., Hoon, M. A., Ryba, N. J. P. & Zuker, C. S. The receptors and

cells for mammalian taste. Nature 444, 288–294 (2006).

119. Treesukosol, Y. & Spector, A. C. Orosensory detection of sucrose, maltose, and

glucose is severely impaired in mice lacking T1R2 or T1R3, but Polycose

sensitivity remains relatively normal. Am. J. Physiol. Regul. Integr. Comp. Physiol.

303, R218–235 (2012).

120. Treesukosol, Y., Blonde, G. D. & Spector, A. C. T1R2 and T1R3 subunits are

individually unnecessary for normal affective licking responses to Polycose:

implications for saccharide taste receptors in mice. Am. J. Physiol. Regul. Integr.

Comp. Physiol. 296, R855–865 (2009).

121. Damak, S. Detection of Sweet and Umami Taste in the Absence of Taste Receptor

T1r3. Science 301, 850–853 (2003).

122. Ninomiya, Y., Imoto, T. & Sugimura, T. Sweet taste responses of mouse chorda

tympani neurons: existence of gurmarin-sensitive and -insensitive receptor

components. J. Neurophysiol. 81, 3087–3091 (1999).

123. Delay, E. R., Hernandez, N. P., Bromley, K. & Margolskee, R. F. Sucrose and

Monosodium Glutamate Taste Thresholds and Discrimination Ability of T1R3

Knockout Mice. Chem. Senses 31, 351–357 (2006).

124. Iwasaki, K., Kasahara, T. & Sato, M. Gustatory effectiveness of amino acids in

mice: Behavioral and neurophysiological studies. Physiol. Behav. 34, 531–542

(1985).

103

Page 111: Gustatory Information Processing By Sam Reiter Thesis

125. Maruyama, Y., Pereira, E., Margolskee, R. F., Chaudhari, N. & Roper, S. D. Umami

Responses in Mouse Taste Cells Indicate More than One Receptor. J. Neurosci. 26,

2227–2234 (2006).

126. Chaudhari, N., Landin, A. M. & Roper, S. D. A metabotropic glutamate receptor

variant functions as a taste receptor. Nat. Neurosci. 3, 113–119 (2000).

127. Chaudhari, N., Pereira, E. & Roper, S. D. Taste receptors for umami: the case for

multiple receptors. Am. J. Clin. Nutr. 90, 738S–742S (2009).

128. Ishimaru, Y. et al. Transient receptor potential family members PKD1L3 and

PKD2L1 form a candidate sour taste receptor. Proc. Natl. Acad. Sci. 103, 12569–

12574 (2006).

129. Nelson, T. M. et al. Taste function in mice with a targeted mutation of the pkd1l3

gene. Chem. Senses 35, 565–577 (2010).

130. Lin, W., Burks, C. A., Hansen, D. R., Kinnamon, S. C. & Gilbertson, T. A. Taste

receptor cells express pH-sensitive leak K+ channels. J. Neurophysiol. 92, 2909–

2919 (2004).

131. Oka, Y., Butnaru, M., von Buchholtz, L., Ryba, N. J. P. & Zuker, C. S. High salt

recruits aversive taste pathways. Nature 494, 472–475 (2013).

132. Caicedo, A., Kim, K.-N. & Roper, S. D. Individual mouse taste cells respond to

multiple chemical stimuli. J. Physiol. 544, 501–509 (2002).

133. Tomchik, S. M., Berg, S., Kim, J. W., Chaudhari, N. & Roper, S. D. Breadth of

Tuning and Taste Coding in Mammalian Taste Buds. J. Neurosci. 27, 10840–10848

(2007).

134. Chandrashekar, J. et al. The Taste of Carbonation. Science 326, 443–445 (2009).

104

Page 112: Gustatory Information Processing By Sam Reiter Thesis

135. Tordoff, M. G., Alarcón, L. K., Valmeki, S. & Jiang, P. T1R3: A human calcium

taste receptor. Sci. Rep. 2, (2012).

136. Tordoff, M. G. et al. Involvement of T1R3 in calcium-magnesium taste. Physiol.

Genomics 34, 338–348 (2008).

137. Galindo-Cuspinera, V., Winnig, M., Bufe, B., Meyerhof, W. & Breslin, P. A. S. A

TAS1R receptor-based explanation of sweet ‘water-taste’. Nature 441, 354–357

(2006).

138. Talavera, K. et al. Heat activation of TRPM5 underlies thermal sensitivity of sweet

taste. Nature 438, 1022–1025 (2005).

139. Cruz, A. & Green, B. G. Thermal stimulation of taste. Nature 403, 889–892 (2000).

140. Cartoni, C. et al. Taste Preference for Fatty Acids Is Mediated by GPR40 and

GPR120. J. Neurosci. 30, 8376–8382 (2010).

141. Caterina, M. J. et al. The capsaicin receptor: a heat-activated ion channel in the pain

pathway. Nature 389, 816–824 (1997).

142. Moon, Y. W., Lee, J.-H., Yoo, S. B. & Jahng, J. W. Capsaicin receptors are

colocalized with sweet/bitter receptors in the taste sensing cells of circumvallate

papillae. Genes Nutr. 5, 251–255 (2010).

143. Bautista, D. M. et al. The menthol receptor TRPM8 is the principal detector of

environmental cold. Nature 448, 204–208 (2007).

144. Abe, J. et al. TRPM8 protein localization in trigeminal ganglion and taste papillae.

Brain Res. Mol. Brain Res. 136, 91–98 (2005).

145. Stocker, R. F. The organization of the chemosensory system in Drosophila

melanogaster: a review. Cell Tissue Res. 275, 3–26 (1994).

105

Page 113: Gustatory Information Processing By Sam Reiter Thesis

146. Dethier, V. G. The Physiology of Insect Senses. (John Wiley & Sons inc.).

147. Liman, E. R., Zhang, Y. V. & Montell, C. Peripheral Coding of Taste. Neuron 81,

984–1000 (2014).

148. Hodgson, E. S., Lettvin, J. Y. & Roeder, K. D. Physiology of a Primary

Chemoreceptor Unit. Science 122, 417–418 (1955).

149. Weiss, L. A., Dahanukar, A., Kwon, J. Y., Banerjee, D. & Carlson, J. R. The

Molecular and Cellular Basis of Bitter Taste in Drosophila. Neuron 69, 258–272

(2011).

150. Marella, S. et al. Imaging Taste Responses in the Fly Brain Reveals a Functional

Map of Taste Category and Behavior. Neuron 49, 285–295 (2006).

151. Thistle, R., Cameron, P., Ghorayshi, A., Dennison, L. & Scott, K. Contact

Chemoreceptors Mediate Male-Male Repulsion and Male-Female Attraction during

Drosophila Courtship. Cell 149, 1140–1151 (2012).

152. Jayaraman, V. & Laurent, G. Evaluating a genetically encoded optical sensor of

neural activity using electrophysiology in intact adult fruit flies. Front. Neural

Circuits 1, 3 (2007).

153. Wang, Z., Singhvi, A., Kong, P. & Scott, K. Taste Representations in the

Drosophila Brain. Cell 117, 981–991 (2004).

154. Raman, B., Joseph, J., Tang, J. & Stopfer, M. Temporally Diverse Firing Patterns in

Olfactory Receptor Neurons Underlie Spatiotemporal Neural Codes for Odors. J.

Neurosci. 30, 1994–2006 (2010).

155. Daly, K. C. & Smith, B. H. Associative olfactory learning in the moth Manduca

sexta. J. Exp. Biol. 203, 2025–2038 (2000).

106

Page 114: Gustatory Information Processing By Sam Reiter Thesis

156. Breslin, P. A., Beauchamp, G. K. & Pugh, E. N. J. Monogeusia for fructose,

glucose, sucrose, and maltose. Percept. Psychophys. 58, 327–341 (1996).

157. Spector, A. C., Markison, S., St John, S. J. & Garcea, M. Sucrose vs. maltose taste

discrimination by rats depends on the input of the seventh cranial nerve. Am. J.

Physiol. - Regul. Integr. Comp. Physiol. 272, R1210–R1218 (1997).

158. Davis, N. T. & Hildebrand, J. G. Neuroanatomy of the sucking pump of the moth,

Manduca sexta (Sphingidae, Lepidoptera). Arthropod Struct. Dev. 35, 15–33 (2006).

159. Christensen, T. & Hildebrand, J. Male-specific, sex pheromone-selective projection

neurons in the antennal lobes of the mothManduca sexta. J. Comp. Physiol. A 160,

553–569 (1987).

160. Pouzat, C., Mazor, O. & Laurent, G. Using noise signature to optimize spike-sorting

and to assess neuronal classification quality. J. Neurosci. Methods 122, 43–57

(2002).

161. Fee, M. S., Mitra, P. P. & Kleinfeld, D. Automatic sorting of multiple unit neuronal

signals in the presence of anisotropic and non-Gaussian variability. J. Neurosci.

Methods 69, 175–188 (1996).

162. Hill, D. N., Mehta, S. B. & Kleinfeld, D. Quality Metrics to Accompany Spike

Sorting of Extracellular Signals. J. Neurosci. 31, 8699–8705 (2011).

163. Perez-Orive, J. Oscillations and Sparsening of Odor Representations in the

Mushroom Body. Science 297, 359–365 (2002).

164. Brillinger. Maximum likelihood analysis of spike trains of interacting nerve cells.

Biol. Cybern. (1988).

107

Page 115: Gustatory Information Processing By Sam Reiter Thesis

165. Kass, R. E. Statistical Issues in the Analysis of Neuronal Data. J. Neurophysiol. 94,

8–25 (2005).

166. Truccolo, W. A Point Process Framework for Relating Neural Spiking Activity to

Spiking History, Neural Ensemble, and Extrinsic Covariate Effects. J. Neurophysiol.

93, 1074–1089 (2004).

167. Butts, D. A., Weng, C., Jin, J., Alonso, J.-M. & Paninski, L. Temporal Precision in

the Visual Pathway through the Interplay of Excitation and Stimulus-Driven

Suppression. J. Neurosci. 31, 11313–11327 (2011).

168. McFarland, J. M., Cui, Y. & Butts, D. A. Inferring Nonlinear Neuronal

Computation Based on Physiologically Plausible Inputs. PLoS Comput. Biol. 9,

e1003143 (2013).

169. Pillow, J. W. et al. Spatio-temporal correlations and visual signalling in a complete

neuronal population. Nature 454, 995–999 (2008).

170. Theunissen, F. et al. Estimating spatio-temporal receptive fields of auditory and

visual neurons from their responses to natural stimuli. Netw. Comput. Neural Syst.

12, 289–316 (2001).

171. Theis, L., Chagas, A. M., Arnstein, D., Schwarz, C. & Bethge, M. Beyond GLMs: A

Generative Mixture Modeling Approach to Neural System Identification. PLoS

Comput. Biol. 9, e1003356 (2013).

172. Martelli, C., Carlson, J. R. & Emonet, T. Intensity Invariant Dynamics and Odor-

Specific Latencies in Olfactory Receptor Neuron Response. J. Neurosci. 33, 6285–

6297 (2013).

108

Page 116: Gustatory Information Processing By Sam Reiter Thesis

173. Tripathy, S. J., Padmanabhan, K., Gerkin, R. C. & Urban, N. N. Intermediate

intrinsic diversity enhances neural population coding. Proc. Natl. Acad. Sci. 110,

8248–8253 (2013).

174. Marr, D. Vision: A Computational Investigation into the Human Representation and

Processing of Visual Information. (W.H. Freeman, 1982).

175. Paninski, L. Maximum likelihood estimation of cascade point-process neural

encoding models. Netw. Comput. Neural Syst. 15, 243–262 (2004).

176. Jortner, R. A., Farivar, S. S. & Laurent, G. A Simple Connectivity Scheme for

Sparse Coding in an Olfactory System. J. Neurosci. 27, 1659–1669 (2007).

177. Churchland, P. Eliminative Materialism and the Propositional Attitudes. J. Philos.

78, 67–90 (1981).

178. Vintch, B., Zaharia, A., Movshon, J. A. & Simoncelli, E. P. Efficient and direct

estimation of a neural subunit model for sensory coding. Neural Inf. Process. Syst.

NIPS (2012).

179. Snitz, K. et al. Predicting Odor Perceptual Similarity from Odor Structure. PLoS

Comput Biol 9, e1003184 (2013).

180. Aldworth, Z. N. Dejittered Spike-Conditioned Stimulus Waveforms Yield Improved

Estimates of Neuronal Feature Selectivity and Spike-Timing Precision of Sensory

Interneurons. J. Neurosci. 25, 5323–5332 (2005).

181. Schwarz, G. Estimating the Dimension of a Model. Ann. Stat. 6, 461–464 (1978).

109