140
Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain by Lisa Margaret D’Alessandro A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Graduate Department of Physiology collaborative with the Institute of Biomaterials & Biomedical Engineering University of Toronto © Copyright by Lisa Margaret D’Alessandro (2015)

Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

Developmental Plasticity

of Tonotopic Maps

in Chinchilla Auditory Midbrain

by

Lisa Margaret D’Alessandro

A thesis submitted in conformity with the requirements

for the degree of Doctor of Philosophy

Graduate Department of Physiology

collaborative with the

Institute of Biomaterials & Biomedical Engineering

University of Toronto

© Copyright by Lisa Margaret D’Alessandro (2015)

Page 2: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

ii

Abstract

Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain

Lisa Margaret D’Alessandro

Doctor of Philosophy

Graduate Department of Physiology

collaborative with the

Institute of Biomaterials & Biomedical Engineering

University of Toronto

2015

Sensory areas of the brain have the remarkable ability to reorganize following changes in

peripheral input, especially during early development. Cortex of altricious species (post-

natal hearing onset) is often investigated. It remains unclear whether neural connections

in auditory midbrain reorganize when precocious subjects (hearing onset in utero; e.g.,

chinchilla laniger) are reared in an enhanced acoustic environment.

Neonatal chinchillas were chronically exposed to a moderately-intense (70 dB

SPL) narrowband (2 ± 0.25 kHz) sound stimulus for 4 weeks. No difference in tone-

specific auditory brainstem response thresholds, hair cell morphology, neural thresholds,

and bandwidth 10 dB above threshold were observed around 2 kHz, suggesting the

exposure stimulus was non-traumatic. Sound frequency maps in central nucleus of

inferior colliculus (IC) were defined with microelectrode recordings. We observed a

significant decrease in the proportion of neurons dedicated to octave bands centered at 2-

and 8- kHz. Changes were not limited to sound-exposure frequencies: increases in low-

frequency representation (below 1 kHz) were observed.

Page 3: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

iii

We describe a c-fos immunolabelling protocol for chinchilla. Bands of neurons

observed following 90-min, 6-kHz stimulation lay ventro-medial to those present

following 90-min, 2-kHz stimulation, consistent with the known tonotopic organization

of IC, and verified herein by electrophysiological recordings. Interestingly, we observed

decreased labeling adjacent to these bands, which we suggest represent inhibitory

regions. Following sound exposure, then 90-min, 2-kHz sound stimulation, the number of

labeled cells both in the 2-kHz region and throughout IC was increased. Sound-exposed

subjects who received no further sound stimulation had c-fos expression patterns similar

to silence controls, suggesting sound-exposure does not alter basal levels of c-fos

expression, but changes the way neurons respond to prolonged (90-min) tone stimulation.

This study contributes to a growing body of literature that suggests that sound

frequency representation in auditory midbrain is altered following passive peripheral

input during development.

Page 4: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

iv

Acknowledgements

I would like to thank my thesis advisor, Dr. Bob Harrison, for the opportunity to

study in the Auditory Science Lab, and for his generosity and support over the years. My

thanks as well to the members of my supervisory committee: Drs. Bill Hutchison, Karen

Gordon, and Willy Wong, who offered many helpful suggestions at all stages of the

research.

I am indebted to my colleagues and fellow members of the Auditory Science Lab

at the Hospital for Sick Children, past and present, for their helpful advice, and to Marvin

Estrada in LAS for assistance developing the surgical technique. I would also like to

thank my Master’s thesis advisor, Dr. Ken Norwich. Many of the valuable lessons I

learned while completing the MSc under his mentorship continued to guide me during the

PhD.

I am grateful for financial support from the Natural Sciences and Engineering

Research Council of Canada (NSERC) and from Ontario Graduate Scholarships (OGS).

The lab is supported by grants from the Canadian Institutes for Health Research (CIHR)

and the Masonic Foundation.

Finally, I would like to thank my family and friends for their enduring support.

Page 5: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

v

To Connie

Page 6: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

vi

Table of Contents

Abstract ............................................................................................................................... ii Acknowledgements ............................................................................................................ iv Table of Contents ............................................................................................................... vi List of Tables ................................................................................................................... viii

List of Figures .................................................................................................................... ix

List of Abbreviations ......................................................................................................... xi

Chapter 1: Introduction ....................................................................................................... 1 1.1 Thesis Overview ................................................................................................... 1

Chapter 2: Background ....................................................................................................... 6

2.1 The Mammalian Auditory System ....................................................................... 6

2.1.1 Auditory periphery: External, middle, inner ear ........................................... 6

2.1.2 Topographic representation along sensory pathways ................................. 12 2.1.3 The auditory nerve ...................................................................................... 14

2.1.4 The central auditory system ........................................................................ 17 2.2 Neural plasticity ................................................................................................. 25

2.2.1 The importance of the early environment on development ........................ 25

2.2.2 Neural plasticity of sensory cortices ........................................................... 26 2.2.3 Cortical vs. subcortical plasticity in neonatal & adult models .................... 28 2.2.4 Mechanisms of neural plasticity ................................................................. 32

2.3 Motivation for undertaking the present research................................................ 37

2.3.1 Choice of the animal model ........................................................................ 38 2.3.2 Effects of an enhanced acoustic environment on IC neurons ..................... 39

2.3.3 C-fos immunohistochemistry for the precocious chinchilla ....................... 40 2.4 Hypothesis and Research Objectives ................................................................. 43

Chapter 3: Experimental Methods & Analysis ................................................................. 44 3.1 Overview ............................................................................................................ 44

3.2 Animal Model .................................................................................................... 46 3.3 Sound exposure stimulus and calibration ........................................................... 47 3.4 Cochlear Function and Structure ........................................................................ 48

3.4.1 Auditory Brainstem Responses ................................................................... 48

Page 7: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

vii

3.4.2 Scanning Electron Microscopy ................................................................... 49

3.5 Electrophysiology methods and analysis ........................................................... 51 3.5.1 Subject preparation ..................................................................................... 51 3.5.2 Surgical technique to access auditory midbrain.......................................... 52

3.5.3 Extracellular recordings .............................................................................. 53 3.5.4 Properties of collicular neurons .................................................................. 56

3.6 c-fos immunohistochemistry methods and analysis ........................................... 58 3.6.1 Development of c-fos histological methods for chinchilla ......................... 58 3.6.2 Subjects ....................................................................................................... 63

3.6.3 Acoustic stimulation to induce bands of labeled neurons ........................... 64 3.6.4 C-fos immunolabeling protocol .................................................................. 64 3.6.5 Image analysis ............................................................................................. 66

Chapter 4 ........................................................................................................................... 69 Results of studies of the cochlea, and of electrophysiological recordings in IC .............. 69

4.1 Cochlear thresholds and hair cell morphology in control and sound-exposed

subjects .......................................................................................................................... 70

4.2 Electrophysiological responses in IC neurons ................................................... 73 4.2.1 Response properties of neurons in central nucleus of IC ............................ 73 4.2.2 Neural representation of sound frequency in IC ......................................... 76

4.2.3 Properties of neural tuning curves in the low-frequency region ................. 81 4.3 Summary ............................................................................................................ 81

Chapter 5: c-fos Immunohistochemistry Results .............................................................. 83 5.1 Tonotopic bands in chinchilla IC ....................................................................... 83

5.2 Electrophysiological data corroborate location of c-fos bands .......................... 85

5.3 Quantification of c-fos data ................................................................................ 86 5.4 Sound-exposed subjects ..................................................................................... 89 5.5 Summary ............................................................................................................ 94

Chapter 6: Discussion ....................................................................................................... 95

6.1 Electrophysiological recordings from neurons in IC ......................................... 96 6.1.1 Changes near the region of the sound-exposure stimulus ........................... 96

6.1.2 Tonotopic map changes in a low-frequency region .................................. 101 6.2 Global neuronal expression patterns: c-fos immunolabeling ........................... 105

6.2.1 Technique development ............................................................................ 105 6.2.2 c-fos expression patterns following sound-exposure ................................ 109

6.3 Methodological considerations ........................................................................ 112 6.4 Conclusions, Future Work................................................................................ 115

References ....................................................................................................................... 117

Page 8: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

viii

List of Tables

Table 3.1: Sample size by experimental group for each of the four experimental

techniques employed ..................................................................................................45

Table 3.2: Primary and secondary antibody dilution combinations tested ........................60

Table 5.1: Average counts of c-fos labelled neurons: tone stimulation vs. silence

controls .......................................................................................................................87

Table 5.2: Average counts of c-fos labelled neurons: sound-exposed subjects vs.

controls .......................................................................................................................89

Page 9: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

ix

List of Figures

Figure 2.1: Anatomy of the auditory periphery ...................................................................7

Figure 2.2: Anatomy of cochlear structures .........................................................................9

Figure 2.3: Hair cell stereocilia, and their imprints on the underside of the tectorial

membrane ...................................................................................................................10

Figure 2.4: Shape and height of outer hair cell stereocilia differ along the length of

the cochlea ..................................................................................................................11

Figure 2.5: Tonotopic organization of the cochlea ............................................................14

Figure 2.6: Frequency tuning curves of auditory nerve fibres in normal subjects,

and those with partial damage to outer hair cell stereocilia .......................................16

Figure 2.7: Major nuclei of the ascending auditory pathway ............................................18

Figure 2.8: Frequency-response areas, spike-time histograms, and their correlated

neuronal types and locations in the cochlear nucleus .................................................19

Figure 2.9: Inferior colliculus: divisions, microanatomy, and connections within

the ascending auditory pathway .................................................................................24

Figure 2.10: Developmental plasticity of monkey striate cortex .......................................27

Figure 2.11: Sound frequency representation in chinchilla inferior colliculus ..................30

Figure 2.12: Cortical over-representation following rearing in an enhanced

acoustic environment ..................................................................................................32

Figure 2.13: Mechanism by which NMDAR-dependent LTP results in the

insertion of AMPA receptors in the post-synaptic membrane ...................................35

Figure 2.14: Schematic of dendritic arborization and synaptic pruning ............................37

Figure 3.1: Overview of experimental design....................................................................46

Figure 3.2: Acoustic spectrum of the neonatal sound exposure stimulus ..........................48

Figure 3.3: Determination of ABR thresholds from an intensity series ............................49

Figure 3.4: Hair cell patterning along the sensory epithelium ...........................................50

Figure 3.5: Chinchilla cochlear frequency-place map .......................................................51

Figure 3.6: Direct electrode placement into IC ..................................................................53

Figure 3.7: Spike discrimination by voltage thresholding .................................................55

Figure 3.8: Electrode track verification .............................................................................56

Figure 3.9: Quantifying characteristics of frequency tuning curves ..................................57

Figure 3.10: Schematic illustrating solution incubations for c-fos ....................................59

Page 10: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

x

Figure 3.11: C-fos expression in a mouse vs. chinchilla model prior to optimization

of duration and temperature of the primary antibody incubation for chinchilla ........61

Figure 3.12: Primary antibody incubation duration and temperature experiments ...........62

Figure 3.13: C-fos specificity controls...............................................................................66

Figure 3.14: Quantification of c-fos immunolabeling results ............................................68

Figure 4.1: Average ABR thresholds for sound-exposed and control subjects .................70

Figure 4.2: Hair cells in the 2-kHz region of the cochlea ..................................................72

Figure 4.3: Minimum thresholds of IC neurons.................................................................74

Figure 4.4: Neural bandwidth 10 dB above threshold .......................................................75

Figure 4.5: Representative tuning curves for low-, mid-, and high-frequency

regions ........................................................................................................................76

Figure 4.6: Characteristic frequency vs. depth for controls and representative

sound-exposed subjects ..............................................................................................78

Figure 4.7: Grouped CF vs. depth plots for all control and all sound-exposed

subjects .......................................................................................................................80

Figure 5.1: Representative sections illustrating c-fos labeling for subjects who

received 90 minutes of pure-tone stimulation vs. silence control ..............................84

Figure 5.2: Electrophysiological recordings corroborate the location of bands

of c-fos labeled neurons present following pure-tone stimulation .............................86

Figure 5.3: Average cell counts vs. frequency position along tonotopic map

for subjects who received pure tone sound stimulation, and silence controls ............88

Figure 5.4: Representative patterning of fos-labeled cells by experimental group ...........90

Figure 5.5: Cell counts for subjects who received 90 min, 2-kHz sound

stimulation, and from those who were sound-exposed, then heard 90 min of

2-kHz sound stimulation ............................................................................................91

Figure 5.6: The width of the 2-kHz band is not different between control and

sound-exposed subjects ............................................................................................. 92

Figure 5.7: Cell counts for subjects who received no sound stimulation, and

for those who heard the sound-exposure stimulus, then received no additional

sound stimulation .......................................................................................................93

Figure 6.1: Schematic of chinchilla frequency tuning curves..........................................102

Figure 6.2: Data from Figures 3 & 5 of Clopton & Winfield (1976) re-plotted ..............103

Figure 6.3: Increased neural representation of frequencies about 1-kHz below

the tone-exposure frequency of 14 kHz (from Oliver et al., 2011) ..........................104

Figure 6.4: Schematic emphasizing the difference in timing between micro-

electrode recordings and c-fos immunohistochemical experiments .........................111

Page 11: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

xi

List of Abbreviations

A1 primary auditory cortex

ABR auditory brainstem response

AMPA α-Amino-3-hydroxy-5-methyl-4-isoxazoleproprionic acid

AMPAR AMPA receptor

arc activity-regulated cytoskeleton

AVCN anteroventral cochlear nucleus

CF characteristic frequency

CN cochlear nucleus

DCN dorsal cochlear nucleus

DNLL dorsal nucleus of the lateral lemniscus

IC inferior colliculus

IEG immediate early gene

IHC inner hair cell

i.p. intraperitoneal

LTP long-term potentiation

LSO lateral superior olive

MET mechanically-gated electrical transduction (channels)

MGB medial geniculate body

MNTB medial nucleus of the trapezoid body

mEPSCs miniature excitatory post-synaptic currents

MSO medial superior olive

MTB medial nucleus of the trapezoid body

NGFI-A nerve growth factor induced-A

NLL nuclei of the lateral lemniscus

NMDA N-methyl-D-aspartate

NMDAR NMDA receptor

OHC outer hair cell

Px post-natal day x; e.g., P28 = post-natal day 28

PVCN posteroventral cochlear nucleus

ROI region of interest

SD standard deviation

SEM scanning electron microscopy; standard error of the mean

SOC superior olivary complex

VNLL ventral nucleus of the lateral lemniscus

Page 12: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

1

Chapter 1

Introduction

In this chapter, we outline the reasoning that led to the undertaking of the current

research. In subsequent chapters, we elaborate on many of the concepts introduced in the

present chapter.

1.1 Thesis Overview

Sensation involves the ability to transduce, encode, and perceive information from

the external environment. A sizeable portion of the brain is devoted to these processes.

Historically, the representation of sensory information in central brain centers was

considered static. However, during the latter half of the 20th century, theoretical and

experimental results began to emerge that provides evidence for a more ‘plastic’ brain.

Plasticity can be broadly defined as the ability for neural connections and function to be

modified depending on sensory input and previous experience.

Page 13: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

2

Acoustic stimuli from the external environment are represented in the brain by

patterns of neural activity established along the sensory epithelium of the cochlea.

Beginning in the cochlea, the representation of sound frequency is organized

systematically (from low- to high-frequency). The resulting neural activity patterns are

transmitted to cortex, maintaining this tonotopic organization. While the representation of

sound within the auditory brain is similar among vertebrates, the maturity of the auditory

system at birth differs between species. Many common laboratory species, such as the

mouse or rat, are altricious species: born with relatively immature auditory systems and

that undergo hearing onset post-natally. Humans, on the other hand, are a relatively

precocious species. At birth, the cochlea is well-developed, and there is some evidence

for hearing responses in utero (e.g., sound-evoked blink startle reflex measured at 24-25

week gestational age by ultrasound imaging, Birnholz and Benacerraf 1983). An

appropriate animal model for studies relating to human auditory development is a

precocious one, such as the chinchilla.

The degree of neural plasticity differs across the lifespan. Following lesions of the

cochlear sensory epithelium, sound frequency representation of the adult auditory cortex

and thalamus are altered. Plasticity of sub-thalamic regions (also in an adult model) are

less pronounced, suggesting that in a mature model, reorganization occurs mainly at the

level of auditory cortex/thalamus in response to peripheral deficits. During early

development, there is evidence for considerable plasticity at cortical as well as sub-

thalamic auditory regions following cochlear lesions. While studies of sensory deficits

provide a useful model of auditory pathway development in a hearing-impaired subject,

they are less instructive as to how auditory pathways develop in a normal-hearing subject.

Page 14: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

3

In contrast to sensory deficits, do acoustically-augmented environments alter neural

representation of sound frequency in a normal hearing subject?

Studies from the Harrison lab suggest that in cat auditory cortex, sound frequency

maps are altered following rearing in an enhanced acoustic environment (Stanton and

Harrison 1996). Later work, also from the Harrison lab, provides evidence that thalamo-

cortical projections are relatively normal in the cat following neonatal cochlear lesions

(Stanton and Harrison 2000), suggesting that reorganization of sound frequency maps

occurs at sub-thalamic auditory nuclei. It is unknown to what extent such changes occur,

if at all, in the inferior colliculus of auditory midbrain (the first sub-thalamic auditory

region) in a precocious animal model following neonatal rearing in an augmented sound

environment.

Our working hypothesis is that the development of neural connections within the

ascending auditory pathway is influenced, in large part, by patterns of sensory activity

elicited by environmental sound stimulation during an early post-natal period.

Specifically, we hypothesize that passive exposure to an enhanced sound environment (in

our case, a chronic, moderate-level narrowband signal centered at 2 kHz) changes the

neural representation of sound frequency in central nucleus of inferior colliculus (IC) of

developing subjects compared with age-matched controls.

To test this hypothesis, neonatal chinchillas (Chinchilla laniger) were exposed for

4 weeks to a moderately-intense (70 dB SPL), narrowband-signal-enriched (2 ± 0.25

kHz) sound environment. We were interested in the structural and functional abilities of

hair cells of the cochlea, following this 4-week period of neonatal sound exposure. The

sound-exposure stimulus was designed to elicit neuronal activation, but not to damage

Page 15: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

4

cochlear hair cells. We conducted pure-tone ABR threshold assessments at frequencies

around the sound-exposure stimulus, and assessed hair cell morphology and distribution

with scanning electron microscopy, also near the 2 kHz region, to examine the functional

and structural integrity of hair cells post-sound-exposure.

We used two techniques to quantify changes in sound frequency representation in

inferior colliculus: 1) in vivo micro-electrode recordings through central nucleus of

inferior colliculus, a direct measure of neural function and a classic technique considered

a “gold standard”; and 2) immunolabeling of the c-fos protein, which is expressed by

auditory neurons after they have been recently “activated” by sound. This technique

provides an indirect histological measure of global neural activation patterns with single-

cell resolution, and makes use of advances in molecular biology. To the best of our

knowledge, there was no previously reported technique for c-fos immunohistochemistry

for the precocious chinchilla, thus our first step here involved developing a c-fos protocol

for the chinchilla. Together, these experiments are designed to provide some of the first

experimental evidence of possible effects of sound exposure on midbrain reorganization

in a precocious species.

Studying the effect of sound on the developmental plasticity of the auditory

system is important for our general understanding of how the auditory system develops.

This work will add to our understanding of subcortical plasticity in an animal model

whose auditory system, in an early developmental epoch, relates closely to that of

humans. While not a direct model, a better understanding of how sound influences the

normal development of auditory pathways may have some relevance for hearing-

impaired children who receive rehabilitative treatment, such as hearing aids or cochlear

Page 16: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

5

implants. These children receive auditory stimulation during a period in which their

auditory structures have significant plasticity.

In the next chapter, we expand on many of the concepts introduced in the present

chapter. We summarize relevant findings from the literature, as it relates to the research

herein: we review the structure and function of the cochlea, and the transmission and

processing of auditory information as it passes to higher auditory nuclei; we review

sensory plasticity in developing and adult models, and introduce mechanisms by which

neural plasticity can occur.

Page 17: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

6

Chapter 2

Background

In this chapter, we review anatomy and physiology of the auditory periphery and central

auditory structures, including inferior colliculus: a “hub” of the auditory system. We

describe the effects of post-natal experience on the development of brain circuitry, and

introduce mechanisms by which neural plasticity can occur. We then discuss the

motivation for undertaking the work, including our choice of animal model and the c-fos

immunolabeling method. The chapter concludes with our research hypothesis and

objectives.

2.1 The Mammalian Auditory System

2.1.1 Auditory periphery: External, middle, inner ear

The three divisions of the ear (external, middle, inner; Figure 2.1) each have

specialized functions. The labyrinth of the inner ear is the organ of hearing and balance.

We focus here on hearing, considering each division of the ear in turn.

Page 18: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

7

In all species, the external ear modifies the acoustic spectrum of environmental

sounds before they reach the tympanic membrane. The pinna and concha filter sound

frequencies, providing cues about the elevation of the sound source. The external

auditory meatus forms a resonant cavity, with a resonant frequency of 3 kHz in humans;

thus, sound frequencies near 3 kHz see the greatest gains. Frequencies in the speech

range (2 - 5 kHz) see gains of 10 - 20 dB (Evans 1982; Moller 1983). The primary

functions of the external ear are to gather, transmit, and localize sounds from the external

environment.

Figure 2.1: Anatomy of the auditory periphery (adapted from Silverthorn, 2004).

Page 19: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

8

The middle ear transforms airborne sound waves into pressure waves in the fluid-

filled cochlea of the inner ear. When sound waves traveling in air (a low-impedance

medium) strike fluid (a higher-impedance medium), almost all of the acoustical energy

(99.7%) is reflected. Two middle ear mechanisms mitigate this impedance mismatch. The

first, and most important, is the concentration of sound energy from the tympanic

membrane, which has a larger surface area, onto the oval window, which has a smaller

surface area (20:1 in humans). The second is the mechanical advantage of the lever action

of the smallest bones in the human body, the middle ear ossicles, resulting from the

length of the malleus being greater than that of the incus (1.4:1 in humans). With these

impedance-matching mechanisms, approximately 60% of sound energy is transmitted to

the cochlea.

The main function of the inner ear is to convert mechanical energy to electrical

energy (nerve impulses). The cochlea (Figure 2.2) is the structure critical to this process:

it houses hair cells, the non-neural auditory receptor cells, within the organ of Corti

(named after the 19th century Italian anatomist Alfonso Corti). The organ of Corti is

attached to the basilar membrane, partially covered by the tectorial membrane (Figure

2.2C). The organ of Corti houses two types of hair cell: a single row of inner hair cells

(IHCs; Figure 2.3C), the sensory receptors that transduce mechanical energy to electrical

energy, and three rows of outer hair cells (OHCs, sometimes referred to as cochlear

amplifiers, Figure 2.3A), which function to amplify low-intensity sounds and to provide

increased frequency resolution and sensitivity. The tallest stereocilia of OHCs are

imbedded in the overlying tectorial membrane (Figure 2.3B), whereas the stereocilia of

IHCs are either entirely free-standing or only loosely attached to the tectorial membrane

Page 20: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

9

(Figure 2.3D). Unlike IHCs, OHCs contain actin filaments, which suggest an innate

ability to contract. One way OHCs are thought to improve frequency discrimination is by

voltage-dependent somatic motility, which locally alters the movement of the basilar

membrane thus giving a more sensitive and sharper frequency response (review:

Fettiplace and Hackney 2006).

Figure 2.2: Anatomy of cochlear structures. (A) Intact cochlea; (B) Cross-section of the

cochlea; (C) Organ of Corti. Adapted from Silverthorn, 2004.

B

C

A

Page 21: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

10

Figure 2.3: Hair cell stereocilia and their imprints on the underside of the tectorial

membrane. (A) Outer hair cell (OHC) stereocilia of a young (post-natal day 29, P29),

normal-hearing chinchilla, seen with SEM. (B) Imprints of individual stereocilia from the

tallest row of OHCs are seen on the underside of the tectorial membrane: OHCs are

attached to the tectorial membrane. (C) Inner hair cells (IHCs) in a young (P29), normal-

hearing chinchilla. (D) “Imprints” from the tallest row of IHC stereocilia are much less

pronounced than those of OHCs (and may in fact be procedural artifacts): IHCs are free-

standing or only loosely connected to the overlying tectorial membrane. (D’Alessandro

2015, unpublished data.)

The shape and stiffness of outer hair cell stereocilia and size of the bundles also

differ depending on their location along the basilar membrane: shorter, stiffer stereocilia

with broader bundles are observed near the cochlear base (Figure 2.4B); at the apex,

stereocilia are more than twice as tall and more flexible with narrower bundles (Figure

A B

C D

Page 22: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

11

2.4A). These variations become important in the present study when we examine mid-

frequency regions of the cochlea that correspond approximately to the frequency of the

sound exposure stimulus.

Figure 2.4: Shape and height of outer hair cell stereocilia differ along the length of the

cochlea. (A) Apical, (B) basal outer hair cell stereocilia from young (4 - 6 week old),

normal-hearing chinchillas. (D’Alessandro 2015, unpublished data.)

Cochlear function is similar to that of a spectrum analyzer, decomposing complex

acoustical waveforms into simpler frequency components. This occurs, in part, due to the

characteristics of the basilar membrane, which change along its length. It is stiff and

narrow near the oval and round windows near the cochlear base; it is more flexible and

wider near the helicotrema at the cochlear apex, as illustrated diagrammatically in Figure

2.5 (p. 14). Pressure waves initiated by the movement of the stapes footplate in the oval

window establish a traveling wave, the envelope of which reaches a maximum at a

position along the basilar membrane that is determined by the frequency of the sound

stimulus (von Békésy, 1960; Figure 2.5, p. 14).

A B

Page 23: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

12

As fluid waves travel through the cochlea, they displace the flexible basilar and

tectorial membranes. Because the pivot points for these membranes differ, displacement

of the basilar membrane causes the tectorial membrane to move across hair cell

stereocilia, creating a shearing force that bends hair cell stereocilia. The tips of hair cell

stereocilia are linked by filamentous structures called tip links that, when stretched, open

mechanically-gated electrical transduction (MET) channels1. Deflection of hair cell

stereocilia towards the tallest stereocilia increases the probability that MET channels are

open. When open, MET channels allow K+ and Ca2+ influx2, depolarizing the cell, and

creating a receptor potential. The receptor potential opens voltage-gated Ca2+ channels on

the hair cell soma. Ca2+ enters the cell, triggering the release of neurotransmitter

(primarily glutamate) from vesicles at the basal end of the hair cell. Neurotransmitter

diffuses across the relatively narrow space between the hair cell and nerve terminals,

binds to receptors on the nerve terminals, and triggers an action potential in the afferent

nerve.

2.1.2 Topographic representation along sensory pathways

The senses are an essential way in which organisms interact with their

environment. A shared feature of most mammalian sensory systems is topographical

representation from sensory cell epithelium to cortex. These sensory cortices are made up

1 To faithfully transduce high-frequency sounds (that have short waveforms), resolution on the order of

microseconds is needed. Since the relatively slow second messenger pathways used in vision and olfaction

would not suffice, a mechanically-gated electrical transduction (MET) channel has evolved.

2 Although hair cells have a high internal concentration of K+, the large electrical gradient (125 mV)

between hair cells and the K+-rich endolymph, the fluid of the scala media, drives K+ ions into the hair cell

through open MET channels.

Page 24: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

13

of orderly representations, or maps, of receptor surfaces. These maps are organized

topographically. That is, neighbouring points of the sensory epithelium are represented

by adjacent locations in the central nervous system. In the visual system, spatial

relationships among retinal ganglion cells are maintained in their central targets, giving

rise to ordered retinotopic maps. Similarly, somatotopic maps exist in the somatosensory

system. In the auditory system, hair cells along the length of the cochlea respond

preferentially to specific sound frequencies. High frequency sounds are detected by hair

cells at the basal end of the cochlea. Low frequency sounds excite hair cells at the apical

end (Figure 2.5). Pure tones stimulate a narrow range of cells along the cochlea; sounds

of increased spectral complexity stimulate a broader region, and the resulting neural

activation patterns are transmitted to central auditory regions of the brain up to cortex.

Given this place coding of sound frequency, we tend to describe neuronal organization

within the auditory system as tonotopic. This tonotopic organization is maintained at all

successive nuclei within the ascending pathway up to cortex, including in the inferior

colliculus, which is the region of interest in the present thesis. Topographic maps are an

efficient way to represent stimulus patterns from the periphery at higher levels in the

central nervous system.

Page 25: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

14

Figure 2.5: Tonotopic organization of the cochlea. High-frequency tones preferentially

activate hair cells at the cochlear base; low-frequency tones, the cochlear apex. (Image

adapted from Purves et al. 1997, after data from von Békésy, 1960)

2.1.3 The auditory nerve

Hair cells are innervated by bipolar neurons of the spiral ganglion, the axons of

which become the auditory portion of the VIIIth cranial nerve, the vestibulocochlear

nerve. Innervation patterns of hair cells differ. Afferent innervation of inner hair cells is

denser than that of outer hair cells: the majority of afferent fibres (90 - 95%), termed

inner radial fibres, innervate inner hair cells. The remaining 5 - 10% (called outer spiral

fibers) innervate outer hair cells. Inner radial fibers become Type 1 (myelinated) cells in

the spiral ganglion; outer spiral fibers become Type 2 (unmyelinated) cells. These cells

are kept separate in the spiral ganglion to preserve cochleotopic organization. Each

afferent innervates a single inner hair cell while each inner hair cell is innervated by

multiple (10 - 20) afferents. Efferents to inner hair cells arise from the uncrossed

Narrow, Thick

Basilar Membrane

Wide, Thin

Basilar Membrane

Page 26: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

15

olivocochlear bundle, while those to outer hair cells arise from the crossed olivocochlear

bundle. Outer hair cells are predominantly innervated by efferent fibres, suggesting these

cells are mainly under control of descending pathways.

Spiral ganglion cells innervate a single inner hair cell, thus each auditory nerve

fiber responds only to a narrow range of frequencies. To determine auditory nerve fibre

responsiveness, recordings are made from individual fibres in response to tone bursts of

varying frequency and intensity. The tip of the resulting isoresponse curve, or (frequency)

tuning curve, indicates characteristic frequency (CF). Frequency tuning curves from eight

neurons originating at different cochlear frequency positions are illustrated in Figure

2.6A. The sharp tuning corresponds to the tuning of the inner hair cell that the auditory

nerve fibre innervates: fibres with low characteristic frequency innervate inner hair cells

at the cochlear apex; high characteristic frequency fibres innervate inner hair cells at the

base of the cochlea. There are many overlapping tuning curves in the auditory nerve

corresponding to the approximately 30,000 (in humans) auditory nerve fibres.

The sharp frequency tuning is vulnerable to cochlear insults such as noise

exposure, ototoxic drugs, and hypoxia. Following partial injury to the stereocilia of outer

hair cells, the sharp tip of the tuning curve remains (depicted as the solid line in the right

panel of Figure 2.6B), yet threshold is elevated by more than 30 dB. The low-frequency

tail becomes hypersensitive, which, together with the threshold elevation, broadens

neural bandwidth (Liberman and Dodds 1984).

Page 27: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

16

Figure 2.6: (A) Frequency tuning curves obtained from individual auditory nerve fibres in

the guinea pig. Adapted from Evans 1972. (B) Following partial damage to the stereocilia

of outer hair cells (depicted semi-schematically in the left panel), tuning curve shape and

sharpness of tuning (solid line, right panel) is altered from normal (dashed line, right

panel). Adapted from Liberman and Dodds, 1984.

B

A

Page 28: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

17

2.1.4 The central auditory system

The primary nuclei of the ascending central auditory system are shown

schematically in Figure 2.7. This figure represents ascending pathways from one ear

only. The cochlear nuclei are the first auditory brainstem nuclei. It is at this level of the

auditory pathway that basic neural response patterns originate, and parallel pathways

emerge (a hallmark of the ascending auditory system). Individual fibers of the cochlear

nerve branch and terminate in one of three divisions of the cochlear nucleus as they enter

the brainstem: the ascending branch innervates the anteroventral cochlear nuclei

(AVCN); the descending branch innervates the posteroventral and dorsal cochlear nuclei

(PVCN and DCN). Tonotopic organization is maintained in all three divisions of the

cochlear nucleus.

Types of neurons and auditory processing vary in the differing divisions of the

cochlear nucleus (Figure 2.8). The ventral divisions of the cochlear nucleus contain

primarily bushy cells (spherical cells, listed in Figure 2.8, are a form of bushy cell),

which signal sound onset, and stellate cells (also called multipolar cells, Figure 2.8),

which signal sustained sound (Oertel et al., 1988). The AVCN functions primarily as a

relay of information from the auditory nerve; neural responses here are similar to those of

the auditory nerve. The cells of the AVCN preserve timing and intensity cues, and project

to both the contra- and ipsi-lateral superior olivary complexes (SOC). Information

transmitted is largely concerned with sound localization mechanisms where interaural

intensity and timing cue differences are compared. The DCN is the first region of the

auditory pathway where lateral inhibition is present (Figure 2.8; bottom 3 categories).

The many cell types here and the complex intrinsic circuitry allows for more complex

Page 29: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

18

processing of auditory signals, thus neural responses in DCN are different from those

seen in the auditory nerve. Cells of the PVCN and the DCN project to the contralateral

nucleus of the lateral lemniscus and inferior colliculus.

Figure 2.7: Major nuclei of the ascending auditory pathway. AVCN: anteroventral

cochlear nucleus; PVCN: posteroventral cochlear nucleus; DCN: dorsal cochlear nucleus;

LSO: lateral superior olive; MSO medial superior olive; MTB: medial nucleus of the

trapezoid body; NLL: nucleus of the lateral lemniscus; IC: inferior colliculus; MGB:

medial geniculate body.

Forebrain

Midbrain

Brainstem

Page 30: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

19

Figure 2.8: Frequency-response areas, spike-time histograms (from which the various

categories are derived), and their correlated neuronal types and locations in the cochlear

nucleus. Solid shading indicates excitation; cross-hatching: inhibition. Note: alternate

cell-type nomenclature, where applicable, is given in parenthesis. Adapted from Evans

1982.

Category Frequency-

Response Area Time Histogram Cell Type Location

Spherical

Build-up

Inhibitory

AVCN/PVCN

PVCN

DCN

Multipolar

(Stellate)

Octopus

Pyramidal

Giant

(Fusiform)

AVCN/PVCN Primary-like

Chopper

Pauser

On

Tone Time

Spikes

Frequency

dB

Frequency

dB Spikes

Tone Time

Page 31: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

20

The superior olivary complex is the first auditory nucleus where information from

both ears meet, thus the emergence of binaural pathways begins here. The main function

of this nuclear complex is sound localization, which is accomplished by different

mechanisms depending on the frequency components of the sound being localized. For

low-frequencies (below approximately 2 kHz in humans), interaural timing differences

are used. Interaural timing differences are computed in the medial superior olive (MSO;

Figure 2.7), which receives binaural excitatory impulses from both ipsi- and contra-

lateral AVCN. These axons vary in length, creating a delay line system. MSO neurons

respond when input from both ears arrives at the same time. This coincidence detection

model (Jeffress 1948) is widely accepted for the localization of low-frequency sounds. In

humans, neural action potentials can phase-lock to frequencies below about 3 kHz (by

contrast, in owls, auditory neurons can phase-lock to sound stimuli up to 9 kHz). This

phase-locking is required for localizing sound based on interaural timing differences. For

sound frequencies above 2 kHz, where phase-locking is not possible, at least in humans, a

different mechanism is used to localize sound: interaural level differences.

In humans, at frequencies greater than about 2 kHz, the head shadow effect

becomes important. When the wavelength of sound frequency is less than the diameter of

the object in its path, the wavelengths are too short to bend around the object, thus

creating an acoustic “shadow”, or region of lower intensity at the far ear. These intensity

differences give an indication of the location of a sound source. Interaural level

differences are computed by neurons of the lateral superior olive (LSO; Figure 2.7) which

receives excitatory projections from the ipsilateral AVCN, and inhibitory (glycinergic)

input from the contralateral ear via interneurons originating in the medial nucleus of the

Page 32: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

21

trapezoid body (MTB; Figure 2.7). Since each LSO encodes sound from the ipsilateral

sound field, input from both LSOs are required to represent the entire sound field in the

horizontal plane. The parallel pathways originating in the cochlear nucleus facilitate these

different mechanisms of sound localization (interaural timing and level differences). The

two pathways for localizing sound are merged in auditory midbrain; both LSO and MSO

project to the inferior colliculus: LSO to the contralateral IC, MSO to the ipsilateral IC.

A distinct set of (monaural) pathways from the cochlear nucleus projects directly

onto contralateral nuclei of the lateral lemniscus (NLL; Figure 2.7), the next nuclear

complex along the ascending auditory pathway. NLL also receives input from ipsilateral

superior olives. Neurons of the NLL have properties and connections similar to those of

IC (Aitkin et al., 1970). It consists of three divisions: dorsal, intermediate, and ventral

(Figure 2.9C; p. 24). The dorsal nucleus is tonotopically organized. A prominent feature

of NLL is neurochemical heterogeneity: approximately 85% of dorsal nucleus neurons

are GABAergic; more than 80% of ventral nucleus neurons are glycinergic; only about

20% of intermediate nucleus neurons are inhibitory (either GABAergic or glycinergic;

Saint Marie et al., 1997). The NLL are not an obligatory synaptic relay; some projections

from cochlear nucleus and SOC terminate directly in IC. The specific role of the NLL in

hearing is unknown. Similar to the superior olives, the output from the NLL terminates in

the IC (Figure 2.7; Figure 2.9C).

The inferior colliculus (IC) is one of the largest auditory nuclei in the mammalian

brain. It has three principle divisions as represented in Figure 2.9A: the central nucleus,

which is exclusively auditory (Aitkin et al., 1994) and essential for normal hearing

(Jenkins and Masterton 1982); the dorsal cortex, which receives substantial descending

Page 33: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

22

projections (e.g., Winer 2005; Winer et al., 1998); and the lateral nucleus, which is multi-

sensory, and integrates auditory and somatic sensory information. To differentiate

between divisions, tuning curves and neural latencies differ. In central nucleus, tuning

curves are sharply tuned, while they are broader in other divisions of IC; also latencies

are shorter for central nucleus neurons vs. those of the other divisions (Syka et al., 2000).

In central nucleus, disc-shaped neurons (called flat cells in rat; Malmierca et al.,

1993) align to form frequency-band laminae (Figure 2.9B). Most laminae run from dorso-

lateral to ventro-medial. This laminar organization is the basis for tonotopic organization

in central nucleus of IC: low frequencies are represented dorso-laterally, high frequencies

ventro-medially (Merzenich & Reid, 1974; Schreiner and Langner, 1997). The axons and

dendrites of stellate cells interact with multiple laminae (Figure 2.9B).

Inferior colliculus is an important ‘hub’ of the auditory system. It is a major site

of convergence, receiving projections from almost all lower-level (binaural and

monaural) nuclei, as well as descending projections from all auditory cortical areas

(Winer et al., 1998; Figure 2.9C). It is an obligatory relay for ascending auditory

pathways to thalamus; neurons originating in IC are the principal source of innervation to

the medial geniculate body (MGB): they make both excitatory (presumed glutamatergic)

and inhibitory (GABAergic; Winer et al., 1996) projections bilaterally to MGB. Thus, IC

neurons are, indirectly, the principle source of innervation to cortex. In addition to these

extrinsic connections, IC also has substantial intracollicular connections (Malmierca et

al., 2003; 2005). Given its many connections, IC is an important link within the auditory

system, processing and integrating information from multiple sources and determining

the form auditory information takes as it is delivered to forebrain regions. It is plausible

Page 34: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

23

that there can be significant neuroplasticity at this nucleus, and this, indeed, is the topic of

the present thesis.

The medial geniculate body (MGB) receives most of its inputs from ipsilateral IC;

projections from MGB are almost entirely ascending and ipsilateral (Figure 2.7). The

degree of descending input to the MBG from cortex is greater than the ascending input

received from IC, implying tight coupling between cortex and thalamus. MGB is divided

into ventral, medial, and dorsal regions. Neurons in the ventral division of MGB have

response properties similar to IC, its main source of input. The tonotopic arrangement of

sound frequency is maintained in the ventral region (Imig and Morel 1984); neither of the

other regions appears tonotopically arranged. Dorsal MGB receives most of its input

from dorsal cortex of IC (Calford and Aitkin 1983). The dearth of intrathalamic

connections to link the three divisions suggests the independence of these nuclei,

compared with the rich intrinsic connections of the IC and cochlear nucleus. Auditory

thalamus has monosynaptic projections to subcortical limbic regions such as the

amygdala (LeDoux et al. 1985; Shinonaga et al. 1994), suggesting the role of MGB may

be more than strictly auditory.

Page 35: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

24

Figure 2.9: (A) Three divisions of the inferior colliculus from Golgi-impregnated material of

the cat. CIC: central nucleus; DIC: dorsal nucleus; EIC: external nucleus of inferior

colliculus. Scale bar indicates 0.5 mm. Adapted from Morest and Oliver 1984. (B) Schematic

of the microanatomy of central nucleus of IC. D: disc-shaped neurons; S: stellate neurons.

Adapted from Oliver et al., 1994. (C) The inferior colliculus (indicated by blue arrows)

receives projections from almost all lower-level nuclei, and is the principle source of

innervation to higher levels of the ascending auditory pathway. Adapted from Pollak et al.,

2003. See text above or List of Abbreviations for full names.

A B

C

D

L

D

L

Efferents

to MGB

Afferents

from:

Contralateral CN, LSO, DNLL;

Ipsilateral VNLL, CN, MSO, LSO

Page 36: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

25

Auditory cortex, located in superior temporal gyrus of the temporal lobe, is where

auditory information interfaces with higher-order communication and cognitive

networks. It has a number of subdivisions. A broad distinction can be made between core

and belt auditory cortical regions. The core auditory cortex is considered the cortical

entry point for the ascending auditory pathway. It is characterized by strong tonotopic

organization and relatively short neural response latencies (Harel et al. 2000). Based on

these features, primary auditory cortex (AI), among other areas, is considered part of core

auditory cortex. Belt areas are less clearly delineated. They receive diffuse projections

from belt regions of MGB, and therefore have a less precise tonotopic arrangement.

Subcortical input to auditory cortex is primarily from ipsilateral thalamus (Figure 2.7).

Primary auditory cortex receives point-to-point connections from the ventral region of

MGB and is tonotopically organized. The grey matter of A1 has six layers that establish

patterns of neural connections with other regions of the brain. For example, layers III and

IV receive input from ventral MGB; layers I and VI receive input from medial MGB

(Huang and Winer 2000). Layers V and VI project to MGB and IC, respectively. Layers

are linked vertically by interneurons (Prieto and Winer 1999) or by specific groups of

pyramidal cells (Ojima et al. 1992).

2.2 Neural plasticity

2.2.1 The importance of the early environment on development

Our sensory systems have evolved to provide central representations of the

external (and internal) environment. During development, when there are normal patterns

of sensory stimulation, sensory pathways develop normally. However, abnormal patterns

Page 37: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

26

of neuronal activity during the maturation process result in the abnormal development of

central sensory pathways. Perhaps one of the earliest examples of the way in which

sensory experience during an early developmental period can “shape” brain activity is by

Konrad Lorenz in the 1930s. He observed that in the absence of a mother goose, newly-

hatched goslings would follow a human being (himself!) as if he were the parent, a

phenomenon termed filial imprinting. This behaviour demonstrates the enormous impact

that the early environment has on brain development.

2.2.2 Neural plasticity of sensory cortices

Numerous experiments over the past few decades have shown that sensory areas

of cerebral cortex can be “re-programmed” as a result of peripheral lesions (reviewed in

Lomber and Eggermont 2006). Neuronal connections can reorganize and synaptic

strength can be modified such that changes in peripheral input result in changes in central

representations of the periphery. These adaptations can be broadly termed: plasticity. For

example, seminal work by Wiesel and Hubel (Wiesel and Hubel 1963; Wiesel and Hubel

1965; Wiesel 1982) demonstrated changes in the functional organization of monkey

striate cortex. In conditions of monocular deprivation (for 18 months, from 2 weeks of

age), the majority of cortical neurons were driven by the non-deprived eye (Figure 2.10).

Subsequent work by Rasmusson (1982) provides evidence for reorganization in primary

somatosensory cortex following digit removal. Sixteen weeks following amputation of

the fifth digit in a raccoon model, the location of receptive fields suggested that the area

of cortex that would normally represent the fifth digit was taken over by input from the

fourth digit. These studies collectively show that areas of the cortex that no longer

receive sensory input become re-wired to process sensory information from areas

Page 38: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

27

adjacent to the lesion in the case of somatosensory studies, or from the non-deprived eye,

in the vision experiments outlined above. Similar results have also been reported in

auditory cortex following partial cochlear lesions. Thirty-five to 81 days after unilateral

mechanical damage to the organ of Corti, the region of contralateral auditory cortex that

would have normally represented frequencies in the range of the lesion, was occupied by

an increased representation of sound frequencies adjacent to the lesioned frequency range

(Robertson and Irvine 1989).

Figure 2.10: Developmental plasticity of monkey striate cortex. Ocular dominance

columns from (A) a control subject, and (B) a subject monocularly-deprived for 18

months from 2 weeks of age. Cortical regions that would normally receive input from the

deprived eye now receive projections predominantly from the non-deprived eye. Scale

bar indicates 1 mm. Adapted from Wiesel 1982.

Page 39: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

28

2.2.3 Cortical vs. subcortical plasticity in neonatal & adult models

Relevant to the present thesis is the distinction between adult and developmental

plasticity, since the experiments described herein focus on experimental alterations to the

auditory system during early post-natal development. Experiments cited in the previous

section were carried out in adult (Robertson and Irvine 1989) and developing (Wiesel and

Hubel 1963; Wiesel and Hubel 1965; Wiesel 1982) animals. These results indicate that

cortical neurons in both mature and developing mammals have the capacity to reorganize

when sensory input is restricted (e.g., partial cochlear lesions) or removed (e.g.,

monocular deprivation). In the adult auditory system, there is evidence for altered sound

frequency representation in auditory thalamus. For example, following restricted

unilateral cochlear lesions resulting in severe mid- to high- frequency hearing loss, there

is an expanded representation of lesion-edge frequencies in the ventral region of medial

geniculate body where mid- to high-frequencies would normally be represented (Kamke

et al. 2003). Harrison (2001) reports less extensive sub-thalamic (midbrain)

reorganization following cochlear lesions (Figure 2.11C) than that reported at the cortical

level (Kakigi et al., 2000). There is also little evidence for reorganization of sound

frequency representation in the cochlear nucleus of mature subjects following cochlear

lesions (Kaltenbach et al., 1996; Rajan and Irvine 1998). The results of these studies

suggest that in the mature animal, it is at the level of the cortex and/or thalamus at which

reorganization mainly occurs in response to a peripheral acoustic deficit.

At the level of auditory midbrain, (i.e., inferior colliculus), reorganization of

sound frequency representation following peripheral lesions has been shown to be more

extensive in a developmental model compared with an adult model. By damaging the

Page 40: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

29

basal cochlear sensory epithelium (hair cells), a high-frequency hearing loss was created

in the neonate (shown in Figure 2.11B as the shaded area along cochlear length). Several

months later, single-unit recordings made in inferior colliculus reveal abnormal

development of sound frequency representation in the neonatal animal (Figure 2.11B)

compared with controls (Figure 2.11A). Similar findings of over-representation of

frequency regions corresponding to the edge of the cochlear lesion have also been shown

at the level of the primary auditory cortex (cat: Harrison et al., 1995; chinchilla: Kakigi et

al., 2000). Also included in Figure 2.11C are the results of the same experimental

manipulation carried out in the adult chinchilla. Together, the results of the neonatal and

adult lesion studies suggest that the auditory midbrain has a greater capacity for

reorganization in an early developmental period rather than at a later, more mature stage

of development.

Page 41: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

30

Figure 2.11: Sound frequency representation in chinchilla inferior colliculus in (A)

control subjects, and subjects in which partial cochlear lesions were induced by oto-toxic

drugs (B) as neonates, and (C) as adults, yielding a hearing loss above 8 kHz. Cross-

hatching in (B) indicates an iso-frequency region. Scale bar: 1mm. Adapted from

Harrison 2006.

Experimental manipulations described thus far have created sensory deficits.

Cochlear lesions resulting in partial deafferentation have reduced auditory input. While

these animal models serve as useful models of sensorineural hearing loss, they are

somewhat less informative about the development of auditory pathways in a normal-

A B C

Characteristic Frequency (kHz) Characteristic Frequency (kHz) Characteristic Frequency (kHz)

Control Sound frequency map following basal

cochlear deafferentation

Neonate Adult Juvenile Subject

Ele

ctro

de

excu

rsio

n (

mm

)

Ele

ctro

de

excu

rsio

n (

mm

)

Ele

ctro

de

excu

rsio

n (

mm

)

0

1

2

3

4

0

1

2

3

4

0

1

2

3

4

0.25 0.5 1 2 4 8 16 0.25 0.5 1 2 4 8 16 0.25 0.5 1 2 4 8 16

Cochlear length apex apex base base Cochlear length

Electrode Tracks 0.5k 0.5k 0.5k

1k 1k 1k

2k 2k 2k

4k 4k 4k

8k 8k 8k

16k

10 kHz iso-frequency

region

D

C

Page 42: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

31

hearing subject. Sensory deficits have been shown to alter the development of sound-

frequency representation; do acoustically-enhanced environments have an influence?

This idea was explored at the level of auditory cortex by rearing newborn kittens in a

spectrally-enhanced acoustic environment (Stanton and Harrison 1996). The dominant

sound source was a tonal stimulus centered at 8 kHz (Figure 2.12C). A moderate sound

intensity, approximately 65 dB SPL, was selected so as to not compromise hair cell

integrity (suggested by normal auditory brainstem response audiograms). Following the

3-month tone-rearing period, subjects were returned to a normal acoustic environment,

and sound frequency representation was assessed at 1 year. A larger area of auditory

cortex was found to be dedicated to the 8-16 kHz octave band (Figure 2.12B, cross-

hatching) compared with controls (Figure 2.12A, cross-hatching).

Page 43: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

32

Figure 2.12: (A) Normal representation of sound frequency. (B) Cortical over-

representation of the 8-16 kHz octave band in a kitten reared in an enhanced acoustic

environment. Cross-hatching in both (A) and (B) indicates the 8-16 kHz octave band

region. (C) Spectrum of the sound-exposure (conditioning) stimulus. sf: sylvian fissure;

aef: anterior ectosylvian fissure; pef: posterior ectosylvian fissure. Data after Stanton and

Harrison 1996.

2.2.4 Mechanisms of neural plasticity

How do patterns of neuronal activity – whether spontaneous or driven by sensory

experience – modify neural circuitry? In this section, we consider several mechanisms of

neural plasticity occurring at the level of the synapse, such as long-term potentiation, as

A

B

C

Page 44: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

33

well as those occurring at the level of the neuron, such as axon elaboration, and synapse

elimination. We begin with a brief overview of Hebb’s postulate.

In 1949, the Canadian psychologist Donald Hebb put forth a number of ideas

relating to the neural basis of learning (a form of neural plasticity). His main thesis,

which has come to be known as Hebb’s postulate, is that, “When an axon of cell A is near

enough to excite cell B and repeatedly or persistently takes part in firing it, some growth

process or metabolic change takes place in one or both cells such that A's efficiency, as

one of the cells firing B, is increased” (Hebb 1949, p. 50). This phrase is known

colloquially as, “cells that fire together, wire together; cells that fire out-of-sync lose their

link”. This casual summary does not, however, distinguish between important temporal

correlations between action potentials of pre- and post-synaptic cells. Repeated arrival of

presynaptic spikes milliseconds before postsynaptic spikes, in many synapse types, leads

to long-term potentiation (spike-timing-dependent plasticity, Bi and Poo 1998; review:

Caporale and Dan 2008).

Some of the first evidence for Hebb’s postulate is the discovery of long-term

potentiation (LTP), first reported by Bliss and Lomo (1973). Synapses have the ability to

undergo activity-dependent modifications in synaptic strength, called synaptic plasticity.

LTP is one of the most well-studied examples of prolonged change in synaptic strength.

It has been used as a model to study the synaptic basis of Hebbian plasticity. While the

capacity for neural plasticity is especially prominent during development, LTP is not

restricted to the developing central nervous system; it has also been reported in adult

models. In inferior colliculus of auditory midbrain, there is some evidence that AMPA

Page 45: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

34

and NMDA receptors mediate excitation (Ma et al., 2002) and that LTP can occur here

(Hosomi et al., 1995, Zhang and Wu 2000).

In the mammalian brain, one of the best understood forms of synaptic plasticity is

N-methyl-D-aspartate-receptor- (NMDAR-) dependent LTP (Figure 2.13). This form of

LTP requires the activation of NMDARs. NMDARs are fast, ligand-gated, ionotropic

receptors. They are activated by coincident presynaptic release of glutamate, the main

excitatory neurotransmitter in the brain, and significant depolarization of the post-

synaptic membrane, which relieves the voltage-dependent Mg2+ block of the NMDAR.

With the Mg2+ block removed, NMDARs are permeable to Ca2+ and monovalent cations

such as K+ and Na+. Ca2+ is an important second messenger critical to the induction of

LTP. The increased concentration of Ca2+ in post-synaptic dendritic spines triggers

intracellular signaling cascades that results in the insertion of AMPA receptors in the

post-synaptic membrane (Figure 2.13). This is the mechanism primarily responsible for

the increase in synaptic strength during NMDAR-dependent LTP. The application of

NMDA antagonists prevents LTP. There is also some evidence to suggest that

morphological changes accompany LTP, such as enlargement of dendritic spines and

post-synaptic densities (review: Yuste and Bonhoeffer 2001).

Page 46: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

35

Figure 2.13: Mechanism by which NMDAR-dependent LTP results in the insertion of

AMPA receptors (AMPARs) in the post-synaptic membrane. See text for description.

Glu: Glutamate; CaMKII: Ca2+/calmodulin-dependent protein kinase II. Adapted from

Kauer and Malenka 2007.

In the auditory system, work by Knudsen and colleagues has revealed a

mechanism of neuronal plasticity in the barn owl. Previous research from this lab

investigated the tuning of neurons to binaural sound localization cues in subjects reared

with altered input during an early post-natal period (Brainard and Knudsen 1993). Baby

owls were reared from P14 to P18 (just as their eyes were opening) for at least 150 days,

with prismatic glasses. This alteration shifts the visual field laterally such that there is a

mismatch between auditory and visual maps in the optic tectum3. The representation of

interaural timing differences is shifted in the direction of the leading ear in both the optic

tectum and the external nucleus of IC, but not in central nucleus of IC, establishing the

3 The optic tectum is the homologue of the superior colliculus in avian species.

Page 47: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

36

external nucleus of IC as a site of plasticity4 (Brainard and Knudsen 1993). A subsequent

study provides evidence for axon elaboration (depicted schematically in Figure 2.14A) as

a mechanism of plasticity that accompanies the functional changes reported in Brainard

and Knudsen (1993). Axons were labeled from their source (central nucleus of IC) to

their point of termination (external nucleus of IC) using an anterograde tracer. In prism-

reared subjects, projection fields were denser and broader compared with controls

(DeBello et al., 2001). Thus, in addition to changes that occur at the level of the synapse,

described in the preceding paragraphs, neural plasticity can occur at the level of the

neuron.

Another mechanism of neural plasticity – at the level of the neuron – is illustrated

in a study of filial imprinting in domestic fowl (White Leghorn; Wallhausser and Scheich

1987). Chicks were imprinted to an acoustic stimulus (400 Hz, 3 bursts/s, presented

continuously at 80 dB SPL) beginning on the 14th day of incubation until hatching.

Subjects were considered imprinted when they moved towards the imprinting stimulus in

an approach test, and when they preferred the imprinting stimulus to a new stimulus in a

simultaneous discrimination task. The Golgi-Cox method was used to study neuron

morphology. Analysis of dendritic segments in specific brain areas 7 days after hatching

revealed a significant decrease in the frequency of dendritic spines in MNH neurons

(mediorostral neostriatum/hyperstriatum). It was hypothesized that this spine loss

(illustrated pictorially in Figure 2.14B) is a critical mechanism of filial imprinting.

4 In the barn owl, interaural timing differences from frequency-specific channels ascend from auditory

brainstem to central nucleus of IC. This information then converges on external nucleus of IC, and is

subsequently conveyed to optic tectum.

Page 48: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

37

Figure 2.14: Schematic of (A) axon elaboration, and (B) synapse elimination,

mechanisms by which plasticity has been reported at the level of the neuron in

developing models. Adapted from Knudsen 2004.

2.3 Motivation for undertaking the present research

As outlined above in Section 2.2.2: Neural plasticity of sensory cortices, many

studies describing sensory system plasticity have revealed alterations in cortical

representation. However, one question that arises is whether cortical plasticity is

intrinsically cortical, or whether it reflects, wholly or partially, reorganization at lower

levels. One of the first steps to address this question is to study whether neurons of sub-

cortical nuclei do indeed reorganize subsequent to altered peripheral input. Relatively

normal thalamo-cortical projection patterns were observed following neonatal deafening

(Stanton and Harrison 2000), suggesting that reorganization occurs at sub-thalamic

levels. At the first sub-thalamic nucleus, inferior colliculus of auditory midbrain (Figure

2.7), there is some evidence for neural reorganization in response to sensory deficits

(Harrison et al., 1998). In large part, the animal models in which midbrain plasticity has

been demonstrated are in altricious species where neonatal manipulations are very early

A B

Post-natal

sensory

experience

Post-natal

sensory

experience

Axon elaboration Synapse elimination

Page 49: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

38

in auditory system development. It remains unclear whether an enhanced acoustic

environment has an effect at this level of the auditory brain in a precocious animal model

(hearing onset in utero, compared with post-natal hearing onset of common laboratory

species, such as rat and mouse). In the present thesis, we ask whether midbrain plasticity

occurs in a precocious animal model in response to post-natal exposure to an enhanced

sound environment.

2.3.1 Choice of the animal model

The chinchilla is a commonly used animal model, particularly for studies of

auditory function. As a precocious species, the maturity of its auditory system, at birth,

differs from that of altricious species. Newborn chinchilla’s cochleae are structurally and

functionally mature. Most hair cells appear adult-like 24 hours after birth (Harrison et al.,

1996). Tonotopic maps in primary auditory cortex and secondary auditory cortical fields

are well-ordered and neurons are sharply tuned by P3 (Pienkowski and Harrison 2005a).

In contrast, onset of cochlear function in altricious species occurs days after birth (P12-

P14 in rat; Geal-Dor et al., 1993). For 2-3 weeks following hearing onset, tonotopic maps

are poorly organized and neurons broadly tuned (Chang et al., 2005; Zhang et al., 2001).

Furthermore, unlike other laboratory species (cat, rat, guinea pig), chinchilla audibility

curves more closely resemble those of humans, also a precocious species (Querleu et al.,

1988), across a broad range of frequencies (Heffner and Heffner 1991; Miller 1970). The

state of the chinchilla’s auditory system, at birth, is similar to that of a human, at birth.

For these reasons, we have selected the chinchilla as the animal model for our studies of

auditory neuroplasticity.

Page 50: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

39

2.3.2 Effects of an enhanced acoustic environment on IC neurons

To date, a few studies have examined the effects of an augmented sound

environment on properties of developing collicular neurons and tonotopic maps, with

differing results. One of the earliest studies (Moore and Aitkin 1975) reports no change in

tuning curves or tonotopic organization after exposing newborn kittens to a continuous

pure-tone, 8 hours/day for the first 50 – 75 days of life. Several papers report changes in

rat tonotopic maps, specifically an increase in the proportion of neurons in central

nucleus tuned to a moderately intense (60-70 dB SPL) sound-exposure frequency

following 14-16-hr exposure to 25-ms tone pips from P9 to P28 (Oliver et al., 2011) and

12h/day exposure to continuous pure tones for 3 post-natal weeks (Poon and Chen 1992).

Click-reared subjects (20/sec, 88.5 dB SPL, from P8 to P19-24) exhibit broader tuning

curves and no change in spontaneous activity, response latency, or tonotopic maps (Sanes

and Constantine-Paton 1985).

More recently, Miyakawa et al. (2013) reported a transient narrowing in tuning

curve bandwidth following chronic tone-pip exposure (7.5 kHz, 100-ms pip duration, 6

pips in a train at 6 Hz, 1 train every 2 s, 60 dB SPL, from P9 to P25). Long-lasting

changes in cortical (but not collicular) tonotopic maps using the same sound-stimulation

pattern were observed. A two-tone rearing paradigm (16 + 40 kHz, 80 dB SPL, from P9

to P17, 22-23 hrs/day) revealed large-scale reorganization of tonotopic maps in IC as

seen by MRI (Yu et al., 2007). These studies have all been done in altricious species

(post-natal hearing onset). To the best of our knowledge, the effects of an enhanced

acoustic environment on the development of tonotopic maps in inferior colliculus of a

precocious animal model have not been reported, and form the topic of the present thesis.

Page 51: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

40

2.3.3 C-fos immunohistochemistry for the precocious chinchilla

Immunolabeling of the c-fos protein has been used to study neuronal activity

patterns in several systems of the vertebrate brain (e.g., vision: Beaver et al., 1993;

somatosensation: Filipkowski et al., 2000; olfaction: Sallaz & Jourdan 1993; audition:

Reimer 1993). C-fos is an immediate early gene (IEG), a family of genes characterized

by rapid and transient induction following a variety of cellular stimuli, among them: the

neuronal depolarization that follows sensory stimulation (reviewed in: Herrera &

Robertson 1996; Hughes & Dragunow 1995; Loebrich & Nedivi 2009; Sheng &

Greenberg 1990). IEG expression occurs in the presence of protein synthesis inhibitors

(cyclohexamide/anisomycin), suggesting they do not require de novo protein synthesis or

the activation of other genes prior to their activation (Cochran et al., 1984; Greenberg et

al., 1986). For these reasons, it has been suggested that IEG expression offers the first

genomic response to cell stimulation.

While experimental protocols exist for many altricious species (mouse (e.g.,

Brown and Liu 1995; Lu et al., 2009), rat (e.g., Friauf 1992; Nakamura et al., 2005),

gerbil (e.g., Scheich and Zuschratter 1995), bat (e.g., Qian and Jen 1994)), our lab is the

first (to our knowledge) to establish a protocol for the precocious chinchilla

(D'Alessandro and Harrison 2014). The c-fos immediate early gene is activated following

acoustic stimulation (e.g., Mello & Pinaud 2006). By chromatically labeling its protein

product, c-fos, we obtain an indirect measure of global neural activation patterns elicited

by acoustic stimulation. We developed a reliable, reproducible c-fos protocol in the

chinchilla using commercially-available reagents. Advantages of using c-fos

immunolabeling (vs., e.g., 2-deoxyglucose) are that it is much quicker (several days vs.

Page 52: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

41

several weeks) and also allows the visualization of neural activity with cellular

resolution.

Recall that the inferior colliculus is arranged tonotopically, with low frequencies

represented dorso-laterally, and high frequencies represented ventro-medially. When a

subject is stimulated with a pure-tone of sufficient duration, the resulting c-fos-labeled

neurons in inferior colliculus lie in a band that corresponds to the frequency used to

induce them (D’Alessandro and Harrison 2014). We use this characteristic of c-fos

expression in inferior colliculus to study whether the neural representation of sound

frequency is altered following rearing in the enhanced acoustic environment. We

compare the width and number of c-fos-labeled neurons in the 2-kHz band in inferior

colliculus of subjects who have heard the sound-exposure stimulus for 4 weeks to those

of age-matched control subjects.

In addition to c-fos, a variety of immediate early genes have been used to map

neuronal activation (NGFI-A: nerve growth factor induced-A, a.k.a. zif268, ERG-1,

KROX-24, and ZENK; and arc: activity-related cytoskeleton). Among the most

commonly studied IEGs, c-fos and NGFI-A exert an effect on the cell by encoding

transcription factors that regulate the expression of downstream target genes termed late-

response genes. Like c-fos, the expression of NGFI-A is highly sensitive to membrane

depolarization. NGFI-A has been shown to regulate the expression of genes that encode

pre-synaptic proteins involved in neurotransmitter release (synapsin I: Thiel et al., 1994;

synapsin II: Petersohn et al., 1995; synaptobrevin II: Petersohn and Thiel 1996) as well as

genes associated with structural stability (neurofilament: Pospelov et al., 1994). The

NGFI-A transcription factor seems well-positioned to relate changes in cell surface

Page 53: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

42

activation with genomic responses, perhaps as part of experience-dependent neuronal

rewiring (review: Knapska and Kazcmarek, 2004).

Arc and immediate early genes like it (e.g., homer1a) are effector genes; they

encode proteins that directly influence cellular function. The mRNA encoded by arc is

transported rapidly to neuronal dendrites, where local translation occurs (Lyford et al.,

1995; Steward et al., 1998; Steward & Worley 2001a). The selective targeting of arc

mRNA to active synapses has been shown to require NMDA receptor activation (Steward

and Worley 2001b). It has also been suggested that arc interacts with an actin-associated

protein (Lyford et al., 1995). Together, these functions have led researchers to propose

that arc is involved in activity-dependent dendritic reconfiguration (Steward and Worley

2001b). Arc has been widely studied in learning and memory related contexts (review:

Bramham et al., 2008).

We selected c-fos immunohistochemistry from among other immediate early

genes for several reasons. First, basal expression levels of this protein are relatively low,

thus allowing us to observe possible increases in neural expression following neonatal

rearing in an enhanced sound environment. The protein product of the NGFI-A

immediate early gene, for example, is expressed at high basal levels (Kaczmarek and

Chaudhuri 1997). Second, the c-fos protein is localized to the cell nucleus, thus each

instance of c-fos-labeling indicates that that neuron has recently been activated (c.f. arc,

which is localized to neuronal dendrites). It was our goal to quantify changes in the

number of neurons activated by sound before and after our experimental treatment. We

thus elected to use c-fos immunohistochemistry to study global neuronal activation

Page 54: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

43

patterns in inferior colliculus of the precocious chinchilla following neonatal rearing in an

augmented sound environment.

2.4 Hypothesis and Research Objectives

Our working hypothesis is that the development of neuronal network connections

within the ascending auditory pathway is driven, in large part, by environmental sound

stimulation during an early post-natal period, consistent with the idea that the brain is

“programmed” to the prevalent acoustic environment. If environmental stimuli during an

early post-natal period are abnormal, we predict alterations in functional organization.

Specifically, we hypothesize that sound frequency maps in central nucleus of inferior

colliculus are altered in neonatal subjects who hear a moderately-intense narrowband

sound stimulus (centered at 2 kHz) for 4 weeks, compared with age-matched controls.

There are four objectives to this research:

1) To determine whether tonotopic maps recorded electrophysiologically in

central nucleus of inferior colliculus differ in sound-exposed subjects compared with age-

matched controls.

2) To establish a c-fos immunohistochemical protocol for the chinchilla.

3) Using this protocol, to compare whether there are differences in the band of c-

fos labeled neurons present following 90-min of pure-tone sound stimulation between

sound-exposed and control subjects.

4) To verify that the sound-exposure stimulus does not damage the cochlea. We

will examine the structure of the hair cells of the cochlea (using SEM), and their function,

by using auditory brainstem responses, ABRs, to estimate hearing thresholds.

Page 55: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

44

Chapter 3

Experimental Methods & Analysis

3.1 Overview

We hypothesize that enhanced peripheral activation during an early post-natal

period will alter the normal representation of sound frequency at the level of the inferior

colliculus. To test this, neonatal chinchillas were chronically exposed to a moderately-

intense (70 ± 5 dB SPL) narrowband (2 ± 0.25 kHz) acoustic stimulus for a period of at

least 4 weeks. Subsequently, micro-electrode recordings of neural action potentials

(spikes) were used to define tonotopic maps in central nucleus of inferior colliculus.

We also visualized global neural activation patterns in chinchilla IC using c-fos

immunohistochemistry with the goal of detecting any abnormal patterns associated with

neonatal sound exposure. This histological method has been used to reveal patterns of

activated neurons in altricious species such as mouse and rat. However, to the best of our

Page 56: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

45

knowledge, no c-fos protocol had been published for the precocious chinchilla. Thus, we

first optimized a protocol for the chinchilla using commercially-available reagents

(D'Alessandro and Harrison 2014).

The neonatal sound exposure paradigm was designed to provide significant

excitation of cochlear neurons, while avoiding acoustic trauma. To study whether sound

exposure altered cochlear function, frequency-specific auditory brainstem responses

(ABRs) were recorded at frequencies around the sound-exposure stimulus. ABR

thresholds of sound-exposed subjects were compared with (non-exposed) control

subjects. To study whether sound exposure altered cochlear structure, the morphology of

hair cell stereocilia and the pattern of hair cell distribution along the cochlear epithelium

in the region of the sound exposure stimulus were assessed in control and sound-exposed

subjects using scanning electron microscopy (SEM).

An overview of experimental timelines is shown in Figure 3.1, and sample sizes

by experimental group for each experimental technique is shown in Table 3.1 below.

Table 3.1: Sample size by experimental group for each of the four experimental

techniques employed. Con: Control; SE: Sound-exposed.

* 1 coronal slice yields 2 samples

SEM

Con SE

# of subjects 2 2

Micro-electrode recordings

Con SE

# of subjects 6 5

# of electrode tracks 16 20

# of multi-units 426 983

ABRs

Con SE

# of subjects

clicks 50 25

tones 26 15

c-fos immunohistochemistry

No

Sound

2 kHz,

90 min

6 kHz,

90 min

SE + 2 kHz,

90 min

SE + No

Sound

# of subjects 5 5 6 5 5

# of samples* 199 158 247 194 249

Page 57: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

46

Figure 3.1: Timelines for (A) ABR, SEM, and extracellular electrophysiological

experiments, and (B) c-fos immunohistochemistry experiments.

3.2 Animal Model

All procedures were approved under The Hospital for Sick Children Animal Care

Committee protocols, following CCAC guidelines. We selected the chinchilla (chinchilla

laniger) as our animal model for several reasons. Primarily, the chinchilla is a precocious

animal: it experiences hearing onset in utero, similar to humans (c.f. altricious rats, who

experience hearing onset between post-natal days 12 to 14 (Geal-Dor et al., 1993);

reviewed in Chapter 2, Section 2.3.1: Choice of the animal model). Thus, the state of the

chinchilla’s peripheral auditory system is similar at birth to that of humans, allowing

some confidence in cross-species extrapolation to humans. Moreover, several practical

considerations contribute to our choice. It is usually free of middle ear infections (Heffner

and Heffner 1991; Miller 1970) and is fairly easily bred and housed. Essential for our

developmental studies is that we can obtain newborn pups. Also, our laboratory has many

A

B

Sound-Exposure

Sound-Exposure Silence

ABR, then

SEM or

Micro-electrode recordings

Birth

Birth P28 P27

P28

ABR

90 min silence,

or 90 min pulsed pure-tone sound stimulation

c-fos immunohistochemistry

Page 58: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

47

years of experience with this species. We have made extensive recordings in both inferior

colliculus (e.g., Harrison et al., 1998) and auditory cortex (e.g., Brown and Harrison

2010; Brown and Harrison 2011; Pienkowski and Harrison 2005a; Pienkowski and

Harrison 2005b).

3.3 Sound exposure stimulus and calibration

We created a narrow-band sound stimulus centered at 2 kHz (2 ± 0.25 kHz), 20

ms rise/fall time, 500 ms on, 1 s off (Adobe Audition 2.0, San Jose, CA, USA; sound

frequency spectrum show in Figure 3.2). The stimulus was chronically presented (24

hrs/day), free-field (Sony Micro Hi-Fi Component System, CMT BX20i, coupled to Sony

speakers, Model #SS-CBX20, Minato, Tokyo, Japan), for at least 4 weeks (28-33 days),

beginning on post-natal day 0 (P0) or P1. The sound-exposure stimulus was calibrated to

be 70 ± 5 dB SPL at the level of the animals’ ears. Calibration stimuli were 1000 ms in

duration. Measurements were made at multiple locations in each cage. The ambient

sound spectrum measured in the animal housing facility was relatively flat with no other

significant peaks. Subjects did not exhibit abnormal behaviour; they appeared to feed

normally. There was no difference in weight between control and sound-exposed subjects

(reported as mean ± SD; controls: 168.3 ± 35.6 g; sound-exposed: 170.4 ± 33.9 g; p =

0.77, t-test).

Page 59: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

48

Frequency (kHz)

0.1 1 10

Re

lative

Le

ve

l (d

B)

0

20

40

60

80

100

Figure 3.2: Acoustic spectrum of the neonatal sound exposure stimulus (2 ± 0.25 kHz).

3.4 Cochlear Function and Structure

To verify that sound exposure did not damage hair cell function or structure, we

measured auditory thresholds using auditory brainstem-evoked responses, and examined

hair cell morphology, particularly around 2-kHz, using SEM. These experiments were

carried out in anaesthetized subjects (ketamine: 15 mg/kg i.p.; xylazine: 2.5 mg/kg, i.p.)

3.4.1 Auditory Brainstem Responses

Auditory brainstem responses (ABRs) to both broadband noise (47-µs clicks; n =

50 & 25 for controls & sound-exposed, respectively) and pure tones around the sound-

exposure stimulus (4 ms pips, Blackman enveloped, at 1, 1.5, 2, 2.5, 3, 4, and 8 kHz; n =

26 & 15 for controls & sound-exposed, respectively) were recorded (Smart EP,

Intelligent Hearing Systems, Miami, FL, USA). Click data were averaged from 1024

sweeps; tone-pip data from 512 sweeps. Skin needle electrodes were placed in a standard

mastoid – vertex configuration. Stimuli were presented monaurally (right ear) through an

Page 60: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

49

insert earphone (ER-2, Etymotic Research, Elk Grove Village, IL, USA) at a range of

intensities, in 10 dB steps. Threshold was taken as the level at which predominant ABR

peaks were just discernible (asterisks in Figure 3.3). Subjects were included in the study

when click thresholds were less than 30 dB SPL.

Figure 3.3: Determination of ABR thresholds from an intensity series. Threshold was

taken as the level at which predominant peak(s) were just discernible (indicated by

asterisks, *); here, 20 dB SPL.

3.4.2 Scanning Electron Microscopy

Anaesthetized subjects (n = 4; 1 female, 3 males) were transcardially perfused

with saline (0.9%) followed by cold fixative (2.5% glutaraldehyde in sodium cacodylate

buffer, pH 7.4, 4 ºC). Cochleae were removed, slowly perfused with 2-3 mL of fixative,

then incubated in fresh fixative for 2 hrs. Samples were post-fixed for 1.5 hours in 1%

Representative Auditory Brainstem Responses

(ABRs) to 47-µs clicks

Norm

0 1 2 3 4 5 6 7 8 9 10 11 12 ms

70 dB SPL

50 dB SPL

40 dB SPL

30 dB SPL

20 dB SPL

10 dB SPL

0 dB SPL

*

*

*

*

*

- ve

Page 61: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

50

osmium tetroxide, and dehydrated through graded ethanol incubations. After cochleae

were dissected open, samples were critical-point dried, sputter-coated with gold and

cochlear morphology was assessed.

In a normal-hearing subject, hair cells are regularly arrayed in 3 rows of outer hair

cells and 1 row of inner hair cells (white arrowheads in Figure 3.4A). If a subject were to

have elevated hearing thresholds, using SEM, we would detect absent hair cells (red

arrows, Figure 3.4B), and hair cell morphology would be noticeably compromised. It is

this level of qualitative comparison of hair cell distribution and morphology that was

made between control and sound-exposed subjects, particularly near the frequency region

of the sound-exposure stimulus (around 2 kHz; Figure 3.5). N.B.: The hair cells shown in

Figure 3.4B are from a control subject who had ABR thresholds greater than 30 dB SPL,

and hence was not studied further.

A Normal-hearing subject B High-threshold subject

Figure 3.4: (A) Hair cell distribution is regularly-patterned in normal-hearing subjects (3

rows of outer hair cells, 1 row of inner hair cells, indicated by white arrowheads). (B) If a

subject were to have elevated thresholds, using SEM, we would detect absent hair cells

(red arrows), and that the morphology of hair cell stereocilia is noticeably compromised.

Page 62: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

51

Figure 3.5: Chinchilla cochlear frequency-place map. Schematic illustrating the region of

the cochlea corresponding to approximately 2 kHz (black arrow) from which hair cells

were examined by SEM.

3.5 Electrophysiology methods and analysis

Following neonatal sound exposure, one of two methodologies used to examine

possible changes in IC neural responses was micro-electrode recordings of spike activity

in central nucleus of IC.

3.5.1 Subject preparation

Eleven chinchillas (chinchilla laniger; 8 females, 3 males) were obtained from

Roseneath Chinchilla (Roseneath, Ontario, Canada). Subjects were aged between post-

natal day 29 (P29) to P34 (140 – 250 g). Animals were anaesthetized with intraperitoneal

injections of ketamine (15 mg/kg) and xylazine (2.5 mg/kg). Subjects were given a one-

half dose of each approximately every hour to maintain sedation for the duration of data

collection, after which the deeply anaesthetized subjects were decapitated. With this

8k

6k

4k

3k

2k

1.5k

1k

500

250

Page 63: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

52

anaesthetic regime, animals respire spontaneously. The avoidance of artificial respiration

(pumping) was important to maintain brainstem stability required for microelectrode

recordings. Body temperature was monitored with a rectal probe and maintained at 37° C

with a thermostatically-controlled heating blanket.

3.5.2 Surgical technique to access auditory midbrain

The inferior colliculus (IC) is usually accessed surgically in one of two ways:

either the overlying cortex is aspirated, or the cerebellum is removed. We developed a

surgical technique to access IC that preserves the integrity of the overlying cortex and

cerebellum. Following tracheotomy and intubation, the cranium was opened above the

junction of the occipital lobe and the cerebellum, and the dura reflected. Surface vessels

were cauterized. A flattened, surgical-grade compressed sponge (Otocell® Ear Wicks,

Boston Medical Products, Westborough, MA, USA) was gently inserted between

occipital cortex and cerebellum. When moistened, this material expands, creating a gap

between them (indicated by the arrow in Figure 3.6), allowing direct visual access of IC

for electrode placement. Silicone oil was applied to the preparation to prevent

desiccation. This technique was a reliable method of accessing IC.

Page 64: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

53

Figure 3.6: Following surgery, inferior colliculus (IC) is visible for direct electrode

placement (indicated by arrow), while the integrity of cortex (Cx) and cerebellum (Cb)

are maintained. Sagittal section obtained upon experiment completion. Scale bar: 1 cm.

3.5.3 Extracellular recordings

Extracellular micro-electrode recordings were made on a vibration-isolation table

in a sound-attenuating booth (IAC). A remotely-controlled microdrive (MCM Controller

Module, FHC, Bowdoin, ME, USA), which was regularly calibrated, held two or four

high-impedance (2 – 4 MΩ) tungsten microelectrodes. Electrodes were spaced ~0.5 mm

apart and were advanced vertically in 50-µm steps to depths of about 3 mm, the

approximate length of the IC, until auditory-responsive neurons were no longer

encountered. We used 2 or 4 bundled electrodes so as to increase neural recording yield.

Stimuli were generated and recordings were stored and analyzed using Tucker Davis

Technology (Gainesville, FL, USA) hardware (System 3 components) and software

(SigGenRP v. 4.4, BrainWare32 v. 9.19).

Once the IC was visible, electrodes were placed using microdrive coordinates. A

search stimulus (50 ms broadband noise, 70 dB SPL) was used to detect auditory

neurons. When responsive neurons were located, 50-ms pure-tones (¼-octave spaced

Page 65: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

54

from 0.1 - 0.4 kHz, ⅛-octave-spaced between 0.4 and 20 kHz) were used to obtain neural

responses. Tones were presented at 4 levels in 10 dB steps starting from ~0 dB SPL, and

at a rate of 3 – 4 stimuli/s. All stimuli were presented twice, pseudorandomly. Based on

the diameter of an electrode tip (< 1 µm) and axon width (~50 µm), we estimate to have

recorded from a few (1-5) units per electrode at each electrode depth.

Tones were delivered monaurally to the right ear using a high-frequency sound

transducer (Intelligent Hearing Systems, Miami, FL, USA) via a short tube and foam ear

tip. Recordings were made from the contralateral IC. Electrode signals were amplified,

and band-pass filtered (0.3 – 5 kHz). Action potentials were discriminated online using

voltage window thresholding. Voltage thresholds were selected to lie just above and

below the noise floor (red horizontal lines in Figure 3.7). Voltages that crossed both

thresholds, and whose voltages had maximum and minimum peaks not more than 700 µs

apart, were considered to be neural action potentials (spikes).

Page 66: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

55

Figure 3.7: Online spike discrimination from background noise by voltage thresholding.

Action potential voltages must cross both thresholds (indicated by red horizontal lines),

and have maximum and minimum peaks not more than 700 µs apart, to be considered a

spike.

Electrode tracks were confirmed histologically to ensure we were recording from

central nucleus of inferior colliculus. Since there are distinct divisions of IC, each with

differing neural properties, it is important to know from which division we recorded.

Therefore, electrode tracks were confirmed histologically. Upon completion of micro-

electrode recordings, a blood vessel near the occipital cortex of the anaesthetized subject

was ruptured. The blood pooled in the valley between cerebellum and cortex (created to

visually place electrodes in IC, described in Section 3.5.2: Surgical technique to access

auditory midbrain). Micro-electrodes were carefully retracted and advanced, using

Upper voltage threshold

Lower voltage threshold

50-ms tone stimulus

Discriminated spike

Discriminated spike

Real-time (online) spike discrimination paradigm

*

*

Page 67: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

56

Narishige micromanipulators, marking electrode tracks in IC through which the

electrodes traversed during the experiment. Following decapitation, the brain was

removed, and placed in fixative for 48 hrs at 4 ºC. One-hundred-micron slices were cut

with a VibratomeTM. Slices were collected on gelatinized slides, dried overnight,

coverslipped, then imaged with MiraxTM. IC divisions, superimposed on the electrode

tracks in Figure 3.8, were estimated from rodent stereotaxic atlases (rat: Watson and

Paxinos 1982; mouse: Franklin and Paxinos 2008). All electrode tracks measured passed

through central nucleus of IC.

Figure 3.8: Electrode tracks were verified upon experiment completion to ensure they

passed through central nucleus. CIC: central nucleus of inferior colliculus; DIC: dorsal

cortex of inferior colliculus; EIC; external nucleus of inferior colliculus. Scale bar: 500

µm.

3.5.4 Properties of collicular neurons

For each frequency-level combination, we obtained an average count of the

number of action potentials (the height of the bars is proportional to the number of

spikes; Figure 3.9A). From these frequency response areas, we obtained frequency

CIC

DIC

EIC

Page 68: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

57

tuning curves (Figure 3.9B). A “trace” of the tuning curve allows us to measure the

following neural properties: characteristic frequency (CF), threshold, and bandwidth 10

dB above threshold (BW10; Figure 3.9C). These properties were determined for

recordings made from each electrode, at each electrode depth.

Figure 3.9: (A) Sample frequency response area (the height of the bars is proportional to

the number of spikes for that frequency/level combination), and (B) tuning curve for a

single electrode depth. (C) A tuning curve “trace” from which we quantify several neural

properties: characteristic frequency (CF); threshold; and BW10: bandwidth 10 dB above

threshold.

A B

C

Threshold:

8.5 dB

BW10: 0.6 octaves 10 dB

CF: 1600 Hz

Page 69: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

58

3.6 c-fos immunohistochemistry methods and analysis

A second, histological method was used to examine potential differences in global

neural activation patterns in IC following chronic, neonatal exposure to a frequency-

specific sound stimulus. This involved: first optimizing the c-fos methodology for the

precocious chinchilla; then determining the appropriate experimental protocol to obtain

tonotopic bands of c-fos-labeled neurons; and finally, quantifying changes in c-fos

expression patterns following neonatal sound exposure.

3.6.1 Development of c-fos histological methods for chinchilla

To reveal nuclear sites where antigen (the c-fos protein) is present, a series of

incubations in various solutions is required. The c-fos protein is localized to the cell

nucleus (Curran et al., 1984; Roux et al., 1990), thus application of a detergent (Triton X-

100) permeabilizes cell membranes, permitting subsequent reagents to access the cytosol

and nuclear material. The general method involves incubations in primary antibody

(binds to antigen) then secondary antibody (binds to primary antibody) solutions, shown

schematically in Figure 3.10. To prevent non-specific binding of antibody to charged

sites on the tissue surface, samples are incubated in a “blocking” solution. In addition to

the detergent mentioned above, this solution contains charged protein molecules (goat

serum, bovine serum albumin) that bind to available charged sites on the tissue surface,

thus preventing non-specific binding. In the final step, DAB (diaminobenzidine) is

oxidized by peroxidase. Since DAB is a peroxidase marker molecule, one of the first

incubations (not shown in Figure 3.10) is in hydrogen peroxide, to block endogenous

enzyme (peroxidase) activity.

Page 70: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

59

Figure 3.10: Schematic illustrating some of the solutions in which tissue samples are

incubated to visualize the c-fos protein, which is localized to the cell nucleus. To adapt

this protocol for the chinchilla, we began by optimizing antibody dilutions.

Since no c-fos immunohistochemical protocol had previously been reported for

the chinchilla, we began by ascertaining optimal dilutions for primary and secondary

antibodies. Combinations of primary (1°), and secondary (2°) antibody dilutions tested

are shown in Table 3.2. The secondary antibody dilution that produced the (qualitatively)

clearest staining with the least background staining was found to be 1:100. We tested two

primary antibody dilutions with the 1:100 secondary dilution and found that a primary

antibody dilution of 1:500 produced the clearest staining (Figure 3.11). Cell staining with

these dilutions, however, was not as clear as could be obtained in a mouse model (c.f.,

Figure 3.11A & B).

DAB

+

H2O2

Colour precipitate

(Diaminobenzidine)

Optimize

antibody

dilutions

Avidin-Biotinylated peroxidase Complex (ABC)

Biotinylated secondary antibody

Primary antibody

Antigen (Fos protein)

Page 71: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

60

Table 3.2: Primary (1°) and secondary (2°) antibody dilution combinations tested. A

combination of primary antibody dilution of 1:500 with a secondary antibody dilution of

1:100 was found to be optimal.

1° 2°

1° 2°

1° 2°

1:50 1:1000 1:600 1:1000

1:500 1:100

1:100 1:1000 1:1000 1:1000

1:1000 1:100

1:200 1:1000 1:600 1:2000

1:500 1:300

1:300 1:1000 1:1000 1:2000

1:1000 1:300

1:400 1:1000

1:500 1:500

1:500 1:1000

1:1000 1:500

1:600 1:1000

1:700 1:1000

1:800 1:1000

1:1000 1:1000

Page 72: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

61

c-fos expression: Chinchilla

Figure 3.11: c-fos expression in a mouse model (A) is clearer than that in a chinchilla

model (B) following optimal antibody dilutions, suggesting further refinement to the

chinchilla c-fos protocol is needed. Scale bar in (A) is 2 mm, in (B): 3 mm. BC: barrel

cortex; Cx: cerebral cortex; Hipp: hippocampus.

B

A c-fos expression: Mouse

10x

10x

D

R

D

L

BC

Cx

Hipp

Page 73: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

62

We next modified primary antibody incubation duration and temperature. We

began with a temperature and duration that produced optimal results in a mouse model,

then decreased temperature and increased incubation duration. Again, a qualitative

assessment of the results suggests that incubation for 48 hours at 4° C produced the

clearest cell staining with the lowest level of background activity (Figure 3.12).

Cell Background

Experiment Temperature Duration Staining Staining

1 Room temp. 24 hrs Poor High

2 4° C 24 hrs Good Reduced but present

3 4° C 48 hrs Best Low

Figure 3.12: Primary antibody incubation duration and temperature experiments and

results. The best cell staining with the lowest level of background staining was

produced under the conditions reported in the 3rd experimental protocol, indicated by

the red rectangle.

Page 74: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

63

To summarize, a primary antibody dilution of 1:500 and a secondary antibody

dilution of 1:100 were taken as optimal for the chinchilla animal model. Further, a

primary antibody incubation duration of 48 hrs and at 4° C was found to be optimal. We

now report on: the acoustic protocol used to obtain bands of c-fos labeled neurons (that

correspond to the tone-frequency used to induce them); and analysis methods to quantify

changes in the pattern of c-fos expression following post-natal rearing in an enhanced

acoustic environment.

3.6.2 Subjects

Twenty-six chinchillas (chinchilla laniger; 12 female, 14 male) were obtained

either on the day of birth, post-natal day 0 (P0), or on P1 from Roseneath Chinchilla

(Ontario, Canada). Subjects were randomly assigned to one of five experiments: two

control groups, where subjects received either no sound stimulation (n = 5), or heard a 2-

kHz pure-tone for 90 minutes (to induce bands of fos-labeled cells; n = 5), or to one of

two sound-exposed groups, where subjects were reared in an enhanced auditory

environment (described above in Section 3.3 Sound exposure stimulus and calibration)

for 4 weeks, then either heard a 90-min 2-kHz tone (n = 5), or received no further sound

stimulation (n = 5). A fifth group heard a 6-kHz tone for 90 minutes (to induce bands of

fos-labeled cells; n = 6). Subjects were 28 - 35 days old (94 – 225 g). Normal hearing was

verified using auditory brainstem response (ABR) thresholds to broadband, 47-µs click

stimuli (Intelligent Hearing Systems, Smart EP).

Page 75: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

64

3.6.3 Acoustic stimulation to induce bands of labeled neurons

All subjects were kept overnight in a 40-dB sound-attenuating booth (IAC) with

ambient noise levels < 20 dB SPL. To induce bands of fos-labeled cells, some subjects

were then stimulated free-field with gated pure tones (2- or 6-kHz; 100 ms on time, 200

ms off, 5-msec cos2-gated) for 90 minutes. All signals were calibrated to be 60 dB SPL at

the level of the subject’s head (Larson Davis, Model 831, with B&K #4230 reference

signal). Silence controls and sound-exposed subjects who received no further sound

stimulation were treated similarly; however, instead of hearing the 90-min pure tone, they

remained in silence for 90 min. Ten minutes before the end of the 90-min period, subjects

were anaesthetized with ketamine (15 mg/kg i.p.) and xylazine (2.5 mg/kg, i.p.) and

returned to the sound field (or silence).

3.6.4 C-fos immunolabeling protocol

At the end of the 90-min stimulation period (or 90-min period of silence),

anaesthetized subjects were transcardially perfused with physiological saline followed by

cold fixative (4% paraformaldehyde in 0.1 M phosphate buffer, pH 7.4). Brains were

removed and kept in fixative overnight at 4º C. Coronal slices (40 µm) were cut

(VibratomeTM) and treated to visualize c-fos as follows. To block endogenous peroxidase

activity, sections were incubated in hydrogen peroxide (0.3%), and rinsed with phosphate

buffer (0.1 M; pH 7.4). To block non-specific antibody binding, sections were incubated

in a blocking solution, (0.1% bovine serum albumin; 0.2% Triton X-100; 2% goat serum

in 0.1 M phosphate buffer). Sections were immersed in the following solutions, each

followed by buffer rinses: primary antibody (rabbit anti-Fos polyclonal IgG, Santa Cruz

Biotechnology Inc.; diluted 1: 500) for 48 hours at 4º C, secondary antibody (goat anti-

Page 76: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

65

rabbit IgG, Jackson Laboratories Inc.; diluted 1:100) for 1.5 – 2 hours at room

temperature, and avidin-biotin complex (Vector Laboratories Inc., diluted 1:50) for 1.5

hours at room temperature. The diluent for antibody solutions and for the avidin-biotin

complex solution was blocking solution. Sections were incubated in 3, 3'-

diaminobenzidine (DAB, 0.05%) for 8 minutes, and staining revealed by application of

0.001% H2O2. Samples were then rinsed with phosphate buffer, mounted on gelatinized

slides, dried, dehydrated, cleared in xylene, and cover-slipped.

Figure 3.13A shows the results of an experiment in which the primary antibody

was omitted (sections were incubated in fresh “blocking” solution for 2 hours). As

expected, no cell staining was observed; the tissue is visibly void of colour, in marked

contrast to sections that undergo the primary antibody incubation (Figure 3.13B). These

results indicate specificity of reagent binding; when an incubation step is omitted,

immunolabeling is abolished.

Page 77: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

66

Figure 3.13: Specificity controls. (A) When primary antibody is omitted, there is no

substrate for the chromogenic reaction; hence neurons expressing the fos protein are not

labeled. (B) Representative section from the same subject in which primary antibody was

included. The 10x magnifications are raw images, obtained prior to ImageJTM

thresholding. Scale bars indicate 3 mm. Adapted from D’Alessandro and Harrison 2014.

3.6.5 Image analysis

Slides were digitally imaged using Mirax ScanTM, which allows visualization of

samples at a range of magnifications, and saved for further analysis. Using ImageJTM,

thresholds were selected which best captured the pattern of fos-labeled neurons in IC.

Cells were included in counts based on size and circularity. Images presented here (unless

otherwise specified) are originals overlaid with “masks” of labeled neurons detected by

ImageJTM thresholding. To quantify the density of labeled neurons, we defined a region

of interest (ROI) that consists of a grid of discrete, equal-sized columns overlaid on the

centre of rotated collicular slices such that the edges of the ROI lay approximately 200-

A B

10x 10x

Page 78: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

67

400 µm away from the edges of the colliculus (see bottom panel, Figure 3.14). The

columns of the grid are parallel to iso-frequency regions. Using a custom-made macro

created in ImageJTM, we obtained counts of labeled neurons within each column of the

ROI (top panel, Figure 3.14). To construct the scale of the abscissa in the resulting

histograms (Chapter 5: C-fos Immunohistochemistry Results; Figures 5.3, 5.5, 5.6), we

related position along the tonotopic axis to frequency as follows. Since we presented

tonal stimulation at 2 different frequencies, we noted the average column in which the

peak of the 2-kHz band lay in representative slices, and that in which the peak of the 6-

kHz tonotopic band lay. We then built the scale such that the distance between the peaks

(2- and 6-kHz) was maintained. The frequency range of the resulting scale relates well to

the tonotopic map in chinchilla IC measured previously in our laboratory via

electrophysiology (Harrison et al., 1998).

Page 79: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

68

Ce

ll C

ou

nts

0

20

40

60

Figure 3.14: Quantification of c-fos immunolabeling results. Representative sample --

from a subject stimulated for 90 min with a 2-kHz, pulsed pure tone -- rotated, with grid

overlaid to obtain cell counts (described in Section 3.6.5 Image Analysis). Scale bar: 500

µm.

We next present results from our functional (auditory brainstem-evoked

responses) and structural (SEM) analyses of the cochlea, as well as results from

electrophysiological (micro-electrode) studies.

M

D

Page 80: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

69

Chapter 4

Results of studies of the cochlea, and of

electrophysiological recordings in IC

The main experimental manipulation in this thesis was prolonged (4-week) exposure of

neonatal animals to a moderately-intense, narrowband sound stimulus. To verify that

sound-exposure did not induce cochlear threshold elevations or hair cell damage, we

assessed cochlear response thresholds using auditory-evoked brainstem responses, and

imaged hair cell stereocilia using scanning electron microscopy (SEM) post-sound

exposure. These results are described in Section 4.1.

We recorded tone-pip-evoked responses from neurons in inferior colliculus of

sound-exposed subjects. We report on several properties of those neurons, relative to

control subjects, in Section 4.2, as well as on the tonotopic representation of sound

frequency in central nucleus of IC.

Page 81: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

70

4.1 Cochlear thresholds and hair cell morphology in control

and sound-exposed subjects

4.1.1 ABR thresholds

In Figure 4.1, thresholds of frequency-specific auditory brainstem responses

(ABRs) are plotted for control (blue symbols) and sound exposed (red symbols) subjects.

At all tone frequencies measured (1 – 8 kHz), average ABR thresholds of sound-exposed

subjects are not different from those of controls (p = 0.98, ANOVA). These data suggest

that neonatal sound-exposure did not cause temporary or permanent threshold changes.

Tone Frequency (kHz)

1 1.5 2 2.5 3 4 8

Rela

tive

Thre

sh

old

(d

B)

-20

0

20

40

60Control (n = 26)

Sound-Exposed (n = 15)

Figure 4.1: Average ABR thresholds for sound-exposed subjects (red symbols) are

not statistically different from those of controls (blue symbols). Error bars are ± SD.

4.1.2 Cochlear hair cell structure

Hair cells were imaged along the length of the cochlea using SEM. In Figure 4.2,

we show representative images of hair cells in the region of the cochlea corresponding

approximately to the sound-exposure stimulus, near 2 kHz (indicated schematically in

Page 82: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

71

Figure 3.5). The SEM analysis is a qualitative study, in which we examined the cochlear

sensory epithelium for loss of hair cells, and for any disruption of the stereociliar bundle.

Hair cells of all subjects appear in the characteristic pattern of 3 rows of outer hair cells, 1

row of inner hair cells (indicated by white arrowheads in Figure 4.2E). The pattern of hair

cells along the sensory epithelium is similar between control and sound-exposed subjects

(Figure 4.2E). Hair cells of sound-exposed subjects have similar morphology as those of

controls (inner hair cells: cf. Figure 4.2A & B; outer hair cells: cf. Figure 4.2C & D).

That is, we were unable to detect any signs of unusual morphology in hair cells of sound-

exposed subjects. Taken together with ABR results, these data suggest that the enhanced

acoustic environment does not affect cochlear hair cell function or structure.

Page 83: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

72

Control Sound-Exposed

Figure 4.2: Hair cells in the 2-kHz region of the cochlea. Morphology of inner hair cells

(c.f. (A) and (B)) and outer hair cells (c.f. (C) and (D)) are similar between groups. (E):

The pattern of hair cells along the sensory epithelium (3 rows of outer hair cells, 1 row of

inner hair cells, indicated by white arrowheads) is also similar between groups.

A B

C D

E

Page 84: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

73

4.2 Electrophysiological responses in IC neurons

We were interested in whether long-term neonatal sound exposure had an

influence on the development of sound representation in auditory midbrain. In this

regard, we examined response properties as well as the tonotopic distribution of auditory

neurons in central nucleus of IC. We recorded from multiple animals, from multiple

litters.

4.2.1 Response properties of neurons in central nucleus of IC

We characterized several response properties of IC neurons from which we

recorded, namely minimum threshold, and bandwidth 10 dB above threshold (BW10), as

described in Chapter 3: Methods, Section 3.5.4: Properties of collicular neurons, and

illustrated in Figure 3.9. We also qualitatively compared the shape of neural response

areas between control and sound-exposed subjects.

Minimum thresholds of neurons in central nucleus of IC are shown in Figure 4.3.

Data reported are from 426 multi-units sampled from control subjects (Figure 4.3A) and

from 983 multi-units recorded from neonatally sound-exposed subjects (Figure 4.3B).

Near the region of the 2-kHz-centered sound-exposure stimulus (specifically, from 1-3

kHz), there was no difference in neural threshold (reported as average ± standard

deviation; controls: 6.7 ± 9.8 dB; sound-exposed: 6.2 ± 7.3 dB, t(163) = 0.46, p = 0.65).

This finding of similar neural thresholds in sound-exposed and control subjects is

consistent with functional studies of the cochlea (Section 4.1.1: ABR thresholds)

suggesting no ABR threshold differences between groups (Figure 4.1).

Page 85: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

74

Controls Sound-Exposed

0.1 1.0 10.0

Ne

ura

l T

hre

sh

old

(dB

SP

L)

-20

0

20

40

601-3 kHz

Characteristic Frequency (kHz)

0.1 1.0 10.0

Ne

ura

l T

hre

sh

old

(dB

SP

L)

-20

0

20

40

60 1-3 kHz

Characteristic Frequency (kHz)

Figure 4.3: Minimum thresholds of IC neurons. There is no difference in neural thresholds near

the region of the sound exposure stimulus (specifically, from 1 – 3 kHz) in neonatally sound-

exposed subjects (B) compared with controls (A). Results shown are from 426 multi-units in

(A), 983 multi-units in (B).

Neural bandwidths 10 dB above threshold (BW10, measured in octaves) are

plotted as a function of characteristic frequency in Figure 4.4. Data reported are from

422 multi-units sampled from control subjects (Figure 4.4A) and from 968 multi-units

recorded from neonatally sound-exposed subjects (Figure 4.4B). In the region of the

sound-exposure stimulus (from 1 – 3 kHz, as indicated in Figure 4.4A and B), there was

no significant difference in BW10 between groups (reported as average ± standard

deviation; controls: 1.7 ± 1.1 octaves; sound-exposed: 1.7 ± 1.0 octaves, t(175) = 0.02, p

= 0.98).

B A

Page 86: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

75

Controls Sound-Exposed

Characteristic Frequency (kHz)

0.1 1.0 10.0

BW

10 (

octa

ves)

0

2

4

61-3 kHz

Characteristic Frequency (kHz)

0.1 1.0 10.0

BW

10

(octa

ves)

0

2

4

61-3 kHz

Figure 4.4: Neural bandwidth 10 dB above threshold (BW10). There is no difference in BW10

near the region of the sound exposure stimulus (specifically, from 1 – 3 kHz) in neonatally

sound-exposed subjects (B) compared with controls (A). Results shown are from 422 multi-units

in (A), 968 multi-units in (B).

Representative tuning curves from low-, mid- and high-frequency regions are

shown in Figure 4.5, for control (left panels) and sound-exposed (right panels) subjects.

Qualitatively, tuning curve shape and off-frequency levels of activity were not different

between groups over the frequency range from which we recorded.

B A

Page 87: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

76

Figure 4.5: Representative tuning curves for low-, mid-, and high-frequency regions for

control (left panels) and sound-exposed (right panels) subjects. Qualitatively, tuning

curve shape, and off-frequency levels of activity were similar between groups. CF:

characteristic frequency.

4.2.2 Neural representation of sound frequency in IC

The increase in characteristic frequency (CF) with increasing electrode depth

observed in control subjects (Figure 4.6A) is consistent with the literature (Harrison et

al., 1998). We calculated the proportion of the tonotopic map that lay within octave-

wide bands with center frequencies as shown in Figure 4.6 (B, D, and F). CF is fairly

Controls Sound-Exposed

CF = 400 Hz

CF = 8 kHz

CF = 2 kHz

Avg

. Num

ber o

f Sp

ike

s p

er S

we

ep

Leve

l [d

B S

PL]

Leve

l [d

B S

PL]

Leve

l [d

B S

PL]

Frequency [Hz] Frequency [Hz]

Page 88: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

77

evenly distributed over a range of frequencies in control subjects (from 0.25 to 8 kHz;

Figure 4.6B). By contrast, in sound-exposed subjects, neurons with low CFs respond

over a greater electode excursion (indicated by arrows in Figure 4.6C & E; all

electrode tracks for representative subjects are shown), and thus a greater proportion

of the tonotopic map is tuned to neurons with low characteristic frequencies (Figure

4.6D, F).

Page 89: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

78

Controls

Characteristic Frequency (kHz)0.1 1 10

De

pth

(m

m)

0

1

2

3

4

Characteristic Frequency (kHz)0.1

250.2

50.5 1 2 4 8 16

% T

on

oto

pic

Ma

p

0

20

40

60

Sound-Exposed, Pup 22

Characteristic Frequency (kHz)0.1 1 10

De

pth

(m

m)

0

1

2

3

4

Sound-Exposed, Pup 28

Characteristic Frequency (kHz)0.1 1 10

Depth

(m

m)

0

1

2

3

4

Characteristic Frequency (kHz)0.1

250.2

50.5 1 2 4 8 16

% T

on

oto

pic

Ma

p

0

20

40

60

Characteristic Frequency (kHz)0.1

250.2

50.5 1 2 4 8 16

% T

onot

opic

Map

0

20

40

60

Figure 4.6: CF vs. depth for controls (A) and representative sound-exposed subjects

(C, E; all electrode tracks recorded from the subject are shown). (B), (D), and (F):

The fraction of the tonotopic map which lies in octave-wide bands (centre

frequencies as indicated) for corresponding CF vs. depth graphs; (A), (C), and (E),

respectively.

F D

A

B

C E

Page 90: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

79

Results from all electrode tracks from all subjects are shown in Figure 4.7.

For control subjects, we recorded from a total of 426 multi-units from 16 electrode

tracks; for sound-exposed subjects: 983 multi-units from 20 electrode tracks. As a

group, sound-exposed subjects show increased representation of frequencies

beginning an octave below the sound-exposure stimulus, between 1 and 0.1 kHz,

relative to controls (cf. Figure 4.7A & C). In general, there is a shift in neural

representation of sound frequency towards lower frequencies (Figure 4.7E), except at

the octave band centered at 4 kHz, where there is an increase in neural representation

(although not significant) that will be discussed further in Ch. 6: Discussion. The

increase in the proportion of the tonotopic map devoted to the octave band centered

at 125 Hz is significant (t(7) = 2.39, p < 0.05; Figure 4.7E). There is a significant

decrease in the fraction of the tonotopic map devoted to the octave band centered at

2-kHz, near the region of the sound-exposure stimulus (t(23) = 3.19, p < 0.05). The

decrease in the octave band centered at 8 kHz is also significant (t(14) = 2.72, p <

0.05; Figure 4.7E).

Page 91: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

80

Controls

Characteristic Frequency (kHz)0.1 1 10

Depth

(m

m)

0

1

2

3

4

Sound-Exposed

Characteristic Frequency (kHz)0.1 1 10

Dep

th (

mm

)

0

1

2

3

4

Characteristic Frequency (Hz)0.

1250.

25 0.5 1 2 4 8 16

% T

onoto

pic

Ma

p

0

10

20

30

Characteristic Frequency (kHz)0.

1250.

25 0.5 1 2 4 8 16

% T

on

oto

pic

Map

0

10

20

30

Characteristic Frequency (kHz)0.

125

0.25 0.

5 1 2 4 8 16

% T

on

oto

pic

Map

0

10

20

30Controls

Sound-Exposed

* *[ [ *[

Figure 4.7: Grouped CF vs. depth plots show increased representation of low

frequencies beginning an octave below the sound-exposure stimulus for sound-

exposed subjects (B) compared with controls (A). Corresponding histograms:

shown separately for controls (C) and sound-exposed (D); plotted together in (E).

All electrode tracks from all subjects are shown. * p < 0.05

A

D

C

B

E

Page 92: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

81

4.2.3 Properties of neural tuning curves in the low-frequency

region

Here we report on characteristics of neural tuning curves in the low-frequency

region, where, in the previous section, we reported an expanded neural representation in

sound-exposed subjects. Specifically, in the frequency region between 0.1 and 1 kHz,

neural thresholds were lower for sound-exposed subjects (reported as average ± standard

deviation; -4.3 ± 10.3 dB cf. 2.8 ± 11.2 dB for controls; t(311) = 8.0, p < 0.001; Figure

4.3). There was no significant difference in BW10 in this region (reported as average ±

standard deviation; controls: 1.4 ± 0.5 octaves; sound-exposed: 1.4 ± 0.6 octaves; t(416)

= 1.54, p = 0.12; Figure 4.4).

4.3 Summary

In this chapter, we have provided evidence that cochlear function and structure

are not compromised following chronic presentation of the moderately-intense sound

exposure stimulus during the first 4 post-natal weeks of life. Hearing thresholds

(estimated from auditory brainstem evoked responses), hair cell morphology, and the

pattern of hair cells along the sensory epithelium (imaged with SEM) are not different

from age-matched controls. Near the region of the sound exposure stimulus, specifically

between 1 and 3 kHz, electrophysiological data suggest there is no difference in neural

thresholds or bandwidth 10 dB above threshold (BW10). Qualitative comparisons of

tuning curves in low- mid-, and high-frequency regions suggest that tuning curve shape

and off-frequency levels of neural activity are not different between groups. Tonotopic

maps in inferior colliculus are altered following neonatal sound-exposure: there is a

Page 93: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

82

marked increase in neural representation of low frequencies (< 1 kHz), and a decreased

proportion of the tonotopic map devoted to frequencies with octave bands centered at 2-

and 8 kHz. In the low frequency region where we observed increased neural

representation (between 100 Hz and 1 kHz), neural bandwidths 10 dB above threshold

(BW10) were not different; neural thresholds were lower for sound-exposed subjects in

this region.

Next, we present c-fos expression patterns in inferior colliculus: at basal levels;

subsequent to 90-min pure-tone stimulation; and following neonatal rearing in the

narrowband-sound-enhanced acoustic environment.

Page 94: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

83

Chapter 5

c-fos Immunohistochemistry Results

In this chapter, we first present data that suggest that the c-fos immunohistochemical

protocol has been successfully optimized for the precocious chinchilla. Using this

method, we then establish a technique to view bands of c-fos labeled neurons following

pure-tone sound stimulation. The location of these bands is compared with characteristic

frequency/electrode depth results from electrophysiological recordings made in a separate

set of subjects. Finally, we assessed neural activation patterns in inferior colliculus in

subjects that were neonatally sound exposed compared with those of controls.

5.1 Tonotopic bands in chinchilla IC

Bands of fos-labeled neurons are visible following 90-minutes of pure-tone sound

stimulation (indicated by arrows in Figure 5.1A and B). Bands produced with a 6-kHz

sound stimulus lie ventro-medial to bands present following a 2-kHz stimulus, in

Page 95: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

84

accordance with the known tonotopic organization of IC (e.g., Merzenich and Reid

1974). Interestingly, we observed decreased cell labeling adjacent to the tonotopic band,

which, to our knowledge, has not been reported using this technique, and is further

discussed in Chapter 6, Section 6.2: Global neuronal expression patterns: c-fs

immunolabeling. Basal levels of c-fos active neurons with no tonotopic banding are

observed in silence controls as illustrated in Figure 5.1C.

Figure 5.1: Representative sections for a subject who received (A) 90 minutes of 2-

kHz sound stimulation, (B) 90 minutes of 6-kHz sound stimulation, and (C) a silence

control. Arrows indicate bands within the inferior colliculus. Scale bars: 500 µm.

Adapted from D’Alessandro and Harrison 2014.

A B

C

2 kHz 6 kHz D

Silence Control

L

Page 96: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

85

5.2 Electrophysiological data corroborate location of c-fos bands

Next, we verified the position of c-fos banding with functional

electrophysiological recordings of tonotopic maps. In a separate set of subjects, we

determined neuron characteristic frequency vs. electrode depth. Bands of c-fos labeled

neurons observed following 90-min, pure-tone stimulation (black arrows, Figure 5.2A

and C) fall within average minimum and maximum depths (represented by vertical lines,

Figure 5.2A and C) at which we recorded from neurons tuned to 2- and 6 kHz.

Representative tuning curves from neurons with characteristic frequencies of 2- and 6

kHz are shown in Figure 5.2B and D, respectively. These results serve to confirm that a

band of fos-labeled neurons induced by specific tonal sound stimulation corresponds to

activity of auditory neurons tuned to that sound frequency.

Page 97: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

86

Figure 5.2: (A) and (C): Bands of c-fos labeled neurons (indicated by black arrows)

following 90 min of pure tone sound stimulation lie within average minimum and

maximum depths (represented as black lines) at which we recorded from neurons tuned

to 2 kHz and 6 kHz. (B) and (D): Representative tuning curves from neurons tuned to 2-

and 6-kHz, respectively. Immunohistochemical and electrophysiological experiments

were conducted on distinct subjects. Scale bars indicate 500 µm. From D’Alessandro and

Harrison 2014.

5.3 Quantification of c-fos data

To quantify c-fos labeled cell density, a grid aligned along the tonotopic

frequency axis in IC and perpendicular to the labeled band was used as shown in Figure

C

A B

D

90 min, 2-kHz Sound Stimulation

90 min, 6-kHz Sound Stimulation

D

L

Frequency [Hz]

Lev

el [

dB

SP

L]

Avg. # of S

pikes/Sw

eep

Lev

el [

dB

SP

L]

Frequency [Hz]

Page 98: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

87

5.3A (described in Chapter 3, Section 3.6.5 Image Analysis). Cell counts so determined

are shown for two experimental groups and one silence control group in the histograms of

Figure 5.3B, C, and D, respectively. For the 2 kHz frequency band in Figure 5.3B and the

6 kHz band in Figure 5.3C, there is a clear increase in the number of labeled cells

compared with off-frequency areas and silence controls (Figure 5.3D). Pooled data from

the 11 experimental and 5 control animals are presented in Table 5.1. Compared with

controls, there is a significant (40%) increase in labeled cells in the 2-kHz region and a

72% increase in the 6-kHz region.

Table 5.1: Average counts of c-fos labelled neurons. Values reported as average ± SEM.

ROI: region of interest.

Experimental Group

Control

(no sound stim.)

2-kHz, 90-min

sound stim.

6-kHz, 90-min

sound stim.

Region

(5 subjects;

199 samplesa)

(5 subjects;

158 samplesa)

(6 subjects;

247 samplesa)

2-kHz band 76.2 ± 2.8 106.9 ± 3.0* --

6-kHz band 64.9 ± 1.2 -- 111.8 ± 2.4*

Entire ROI 306.1 ± 7.7 333.9 ± 9.4 379.0 ± 9.1

aOne coronal slice yields 2 samples

*Significant (p < 0.001) increase, compared with controls in the same region

Page 99: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

88

Avg. N

um

ber

of C

ells

0

10

20

30

40

50

0.25 0.5 1 2 4 8 16 32

2-kHz bandB

Avg.

Nu

mb

er

of

Cells

0

10

20

30

40

506-kHz bandC

0.25 0.5 1 2 4 8 16 32

Frequency Position Along Tonotopic Axis (kHz)

Avg

. N

um

be

r o

f C

ells

0

10

20

30

40

50D

0.25 0.5 1 2 4 8 16 32

Figure 5.3: (A) Representative sample, rotated, with grid overlaid to obtain cell counts

(see Chapter 3, Methods, Section 3.5.4 Image Analysis). Average cell counts (± SEM) for

subjects who received (B) 90-min 2-kHz sound stimulation, (C) 90 min 6-kHz sound

stimulation, and (D) silence control subjects. Arrows denote regions of decreased cell

labeling. Scale bar in (A) indicates 500 µm. Adapted from D’Alessandro and Harrison

2014.

Silence Control

A

Page 100: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

89

5.4 Sound-exposed subjects

Subjects reared in the enhanced sound environment who then heard a 2-kHz pure

tone for 90 min show an increase in c-fos labeled cells over a broad range of the inferior

colliculus compared with control subjects who heard only the 90-min 2-kHz pure tone

(Figure 5.5A; qualitative comparison: Figure 5.4A & B). The increase in both the number

of c-fos-labeled cells in the 2-kHz band (Figure 5.5B), and over the entire region of

interest (Figure 5.5C) is significant, compared with controls subjects who heard only the

2-kHz tone for 90 min (values listed in Table 5.2). There was no difference in the width

of the 2-kHz band (taken as the distance between the two “troughs”, or regions of

decreased neural labeling; Figure 5.6) between groups.

Table 5.2: Average counts of c-fos labelled neurons (± SEM) by experimental group.

Statistical tests were performed between pairs of groups, as indicated; corresponding p-

values are reported.

Number of cells

Experimental Group 2-kHz band ROI

2 kHz, 90 min 106.9 ± 3.0

p < 0.001

334.0 ± 9.4

p < 0.001 Sound-Exposed

+ 2 kHz, 90 min 184.9 ± 3.8 604.6 ± 10.3

No Sound 76.2 ± 2.8

p = 0.4

306.1 ± 7.7

p = 0.8 Sound-Exposed

+ No Sound 79.2 ± 2.1 309.0 ± 6.3

Page 101: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

90

90 min 2-kHz Sound Stim.

Sound-Exposed + 90 min 2-kHz Sound Stim.

No Sound Stim.

Sound-Exposed + No Sound Stim.

Figure 5.4: Representative patterning of fos-labeled cells, by experimental group.

A

D C

B

Page 102: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

91

Frequency Position Along Tonotopic Axis (kHz)

Avg

. N

um

be

r o

f C

ells

0

20

40

60

80

2-kHz, 90 min

Sound-Exposed + 2-kHz, 90 min

0.25 0.5 1 2 4 8 16 32

2-kHz band

ROI

2-kHz Band

Nu

mb

er

of ce

lls

0

100

200*

2 kHz90 min

Sound-Exposed

+ 2 kHz 90 min

Num

ber

of C

ells

0

200

400

600

ROI

2 kHz90 min

Sound-Exposed

+ 2 kHz 90 min

*

Figure 5.5: (A) Cell counts for subjects who received 2-kHz, 90 min sound stimulation

(red bars; 158 samples from 5 subjects), and from those who were reared in an augmented

acoustic environment, then received 90 min of 2-kHz sound stimulation (blue bars; 194

samples from 5 subjects). One coronal slice produces 2 samples. Error bars show SEM.

Differences in cell counts both in the 2-kHz band (B) and over the entire region of interest

(ROI; C) are significant, * p < 0.001.

B C

A

Page 103: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

92

Figure 5.6: The width of the 2-kHz band – taken as the distance between “troughs”, or

regions of decreased c-fos labeling – is not different between control (A) and sound-

exposed (B) subjects.

Next, we wanted to discern whether this increase in c-fos-labeled neurons is due

to sound-exposure alone. The sound-exposed subjects described in the previous

paragraph have, in effect, received 2 “treatments”: sound-exposure, and the 2-kHz tone

for 90 min. Thus, we conducted a separate set of experiments where subjects were

exposed to the enhanced sound environment, then received no additional sound

stimulation. Quantitatively, there is no difference in the number of labeled neurons in

2-kHz band

A

B

Page 104: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

93

either the region where the 2-kHz band would be (Figure 5.7B) or over the entire region

of interest (Figure 5.7C). These results suggest that there is no effect of sound-exposure

alone on basal levels of c-fos-active cells (cf. Figure 5.4C & D).

Frequency Position Along Tonotopic Axis (kHz)

Avg. N

um

ber

of C

ells

0

20

40

60

80 No Sound Stim.

Sound-Exposed + No Sound Stim.

0.25 0.5 1 2 4 8 16 32

"2-kHz band"

ROI

"2-kHz Band"

Num

ber

of cells

0

100

200

NoSoundStim.

Sound-Exposed

+ No Sound

Stim.

p = 0.4

ROI

Nu

mb

er

of ce

lls

0

200

400

600

NoSoundStim.

Sound-Exposed

+ No Sound

Stim.

p = 0.8

Figure 5.7: (A) Cell counts for subjects who received no sound stimulation (neither the 4-

week sound-exposure stimulus, nor any subsequent pure-tone stimulation; red bars; 199

samples from 5 subjects), and from those who heard the sound-exposure stimulus, then

received no additional sound stimulation (blue bars; 249 samples from 5 subjects). One

coronal slice yields 2 samples. Error bars show SEM. There is no significant difference in

cell counts where the 2-kHz band would be (B), or over the entire region of interest (ROI;

C).

B

A

C

Page 105: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

94

5.5 Summary

In this chapter, we have demonstrated the viability of the c-fos

immunohistochemical protocol for the precocious chinchilla. We reported on resting

levels of c-fos activity in inferior colliculus, and the bands of c-fos-labeled neurons

present following 90 minutes of pure-tone stimulation of differing frequencies. Bands

present following higher-frequency stimulation (6 kHz) lay ventro-medial to those

present following lower-frequency stimulation (2 kHz). These findings are consistent

with the tonotopic organization of central nucleus. Electrophysiological recordings made

in a distinct set of subjects corroborated the location of these bands. Subsequent to the 4-

week neonatal sound exposure period, we observed a significant increase in the number

of c-fos-labeled neurons, both in the 2-kHz band, and over the entire region of interest.

There was no difference in the width of the 2-kHz band. The 4-week sound exposure

period does not affect basal levels of c-fos expression: subjects who heard the sound

exposure stimulus then received no further sound stimulation had c-fos expression

patterns similar to controls. There was no difference in the number of c-fos-labeled cells

either where the 2-kHz band would be, or within the larger region of interest. These data

suggest that sound-exposure has changed the way in which neurons in IC subsequently

respond to prolonged (90-min) periods of pure-tone sound stimulation.

We now discuss the implications of all reported results, and suggest mechanisms

by which they may have occurred.

Page 106: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

95

Chapter 6

Discussion

In this thesis, we sought to examine whether neural activation patterns initiated at the

periphery could influence the normal representation of sound frequency in central

auditory structures. We hypothesized that increased activity during an early post-natal

period would alter representation of sound frequency in inferior colliculus, the first sub-

thalamic auditory nucleus. We discuss evidence from both electrophysiological and

immunohistochemical experiments, both of which support the hypothesis. We present

conclusions, and discuss future work.

Page 107: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

96

6.1 Electrophysiological recordings from neurons in IC

6.1.1 Changes near the region of the sound-exposure stimulus

During development, when there are natural patterns of sensory stimulation,

ascending pathways develop normally. However, unusual patterns of stimulus-driven

neuronal activity can result in the abnormal development of central sensory maps (e.g.,

Miyakawa et al., 2013; Schreiner and Polley 2014). Both sensory deficits and enhanced

sensory environments can be viewed as different-from-normal environments, and hence

result in abnormal central maps. We found a significant decrease in neural representation

around the sound exposure stimulus (2 ± 0.25 kHz) in the inferior colliculus of the

precocious chinchilla. We describe this change in neural representation around 2 kHz as a

form of plasticity and not due to a noise-induced hearing loss: neural thresholds, neural

bandwidth 10 dB above threshold (BW10), ABR thresholds, and hair cell morphology for

sound-exposed subjects in this region were not different from those of controls. We

cannot report, with the experimental measurements used in the present thesis, on the state

of the cochlear nerve.

In a study by Kujawa and Liberman (2009), the authors exposed 16-week-old

mice to an octave band of noise (8-16 kHz) for 2 hours at 100 dB SPL. Level and

duration were adjusted such that measures of cochlear thresholds (among them, ABRs)

were elevated for several days before returning to normal. They report that in regions

where temporary threshold shift was maximal (around 32 kHz), despite evidence of

normal hair cell function, Wave I ABR amplitudes were less than half their pre-exposure

values. In frequency regions where threshold shifts were less pronounced (around 12

kHz), Wave I amplitudes recovered more fully. They also report degeneration of pre- and

Page 108: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

97

post-synaptic elements of the inner hair cell in the basal half of the cochlea (high

frequency regions). These results suggest that although ABR thresholds recover, there

may be downstream effects on auditory nerve anatomy and function. The sound exposure

stimulus used in the present study was presented at a considerably lower level (70 dB

SPL, vs. 100 dB SPL). Nonetheless, the intriguing results reported in Kujawa and

Liberman (2009) suggest that future work involving protracted presentation of a sound

stimulus be accompanied by measurements of Wave I ABR amplitudes to study the

function of cochlear neurons, in addition to techniques that examine hair cell integrity

(ABR thresholds, SEM). With the data presented herein, we can make no assertions about

the integrity of the cochlear nerve.

One plausible mechanism for the decrease in neural representation we observed at

2 kHz may relate to homeostatic decreases in synaptic strength following potentiation of

neurons in the 2-kHz region. Turrigiano and colleagues (Turrigiano et al. 1998) measured

the magnitude of AMPA-mediated synaptic currents as a function of activity in cultured

pyramidal neurons in rat visual cortex. When synaptic activity was reduced by growing

cells for 48 hours in culture containing tetrodotoxin (TTX), a Na+ channel blocker, the

amplitude of miniature excitatory post-synaptic currents (mEPSCs) was significantly

increased. Conversely, enhancing activity by growing cells in culture containing the

GABAA antagonist bicuculline resulted in a significant decrease in the amplitude of

mEPSCs. These results suggest that the amplitude of the mEPSCs depends on prior

activity levels of the neuron. In the present study, the increased neural firing associated

with prolonged exposure to the sound stimulus may have induced a homeostatic decrease

in synaptic strength at those frequencies such that when tones in the region of 2 kHz are

Page 109: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

98

subsequently presented to determine frequency response areas, neural responses recorded

at the level of IC are decreased.

There were several octave bands in the low-frequency region where neural

representation was increased in sound-exposed subjects (125 Hz, 500 Hz). The increase

at 125 Hz was significant. At higher frequencies, sound-exposed subjects had greater

neural representation at the 4-kHz octave band. This difference, however, was not

significant. The putative decrease in synaptic strength around the 2-kHz region may lead

to a release from lateral inhibition (or disinhibition) of neurons neighbouring the sound-

exposure stimulus, thus leading to an increase in representation of frequencies at the

“edges” of the sound exposure stimulus.

In both developing and adult animal models, neural representation of sound

frequency can be altered by exposing animals to an enriched sound environment.

However, methods to induce these changes differ. To induce changes in sound frequency

representation in developing animals, passive exposure to the sound stimulus suffices

(e.g., de Villers-Sidani et al., 2007; Stanton and Harrison 1996). As a result of the pulsed

pure-tone and frequency-modulated stimulation, respectively, cortical representation of

the stimulus frequency increases. Adult animals must be trained to attend to the acoustic

stimulus (e.g., Recanzone et al., 1993) or the stimulus must be paired with electrical

microstimulation of cholinergic or dopaminergic systems (nucleus basalis: Kilgard &

Merzenich 1998; ventral tegmental area: Bao et al. 2001) in order to observe changes in

sound frequency maps. These results suggest that in adult models, acoustic stimuli must

be behaviourally relevant to induce plastic change, at least at the level of the cortex.

Page 110: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

99

Recent studies by Eggermont and colleagues challenge the notion that the mature

auditory system is immune to changes resulting from passive exposure to a sound

stimulus. Passively exposing 75-day old cats to multitone complexes (Norena et al.,

2006) and adult cats to noise bands (Pienkowski & Eggermont 2009, 2010, review:

Pienkowski & Eggermont 2011) result in decreased responses of cortical neurons in A1

to the majority of frequencies comprising the exposure stimulus. The results of these

experiments together with those in the previous paragraph suggest that the structure of

the auditory stimulus may be an important determinant of possible sound-exposure-

induced neural reorganization.

Following this line of thought, another potential mechanism for the decrease in

neural representation at 2 kHz reported in the present thesis may relate to how different

types of auditory stimuli have differing effects on the auditory system. In previous studies

that report increased neural representation of the sound-exposure frequency in IC

following rearing in an enhanced acoustic environment (e.g., Oliver et al., 2011; Poon

and Chen 1992), the authors presented either pulsed or continuous pure tones. In the

present study, we used a narrowband noise stimulus centered at 2 kHz. We report

significant decreases in neural representation of the octave band centered at 2 kHz. In a

study by Merzenich and colleagues (de Villers-Sidani et al., 2008), neonatal rats were

exposed to a 70 dB SPL bandlimited-noise (extending from 5-20 kHz) or notched noise

(extending from 0-5 kHz and 20-30 kHz, with a noiseless notch from 5-20 kHz) from P7

(before hearing onset) to P20. In both cases, the authors report, among other findings, a

significant decrease in frequencies that lay within the noise bands, and marked over-

representation of frequencies outside the noise-exposure regions. Taken together with the

Page 111: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

100

previous IC studies, including the present one, these results suggest that pure tones and

noise are perceived differently by the auditory system. Pure tones strictly stimulate the

auditory system. Noise serves to mask relevant environmental sound inputs, raising the

threshold of the subject in that frequency range, and preventing them from hearing the

normal patterning of sound that would promote normal development of sound frequency

maps there. In clinical audiology for example, noise is used to mask hearing in a “non-

test ear”, effectively raising thresholds in that ear, to test the hearing in the other “test

ear”, much as an optometrist covers one eye while testing vision in the other eye (Katz

and Lezynski 2002). Masking also prevents acoustic cross-over between ears (measured

in a mouse model: Harrison, Negandhi, Allemang, D’Alessandro, & Harrison 2013). In

the study discussed above by Merzenich and colleagues (de Villers-Sidani et al., 2008),

neural responses in cortex were investigated. Our study is the first, of which we know, to

report changes in sound frequency maps in inferior colliculus, a sub-cortical region,

following rearing in an environment enhanced with a moderately-intense narrowband

noise.

We did not observe an increase in neural representation at the sound exposure

frequency. This may relate to the developmental state of the animal model used.

Altricious animal models, such as the rat used in the aforementioned studies, have

auditory systems that are less mature at birth than a precocious species such as the

chinchilla (see Chapter 2: Background, Section 2.3.1: Choice of the animal model).

Results reported here for the chinchilla reflect changes that occur in a more

developmentally-mature auditory system at the time of the experimental treatments, and

thus may differ from those reported in altricious species.

Page 112: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

101

6.1.2 Tonotopic map changes in a low-frequency region

To gain a more accurate impression of the pattern of neural excitation caused by

the neonatal exposure stimulus, it is instructive to consider the frequency response areas

of auditory neurons. Tuning curves elucidate how neurons respond to tones of varying

frequency/level combinations. Rather than having a fixed shape, tuning curve shape

varies with frequency (Figure 6.1). Neurons with low characteristic frequencies (CFs)

tend to have V-shaped tuning curves, while those with high CFs tend to have a steep,

high-frequency cut-off region and a longer low-frequency “tail”. At threshold, the pure-

tone frequency that activates the neuron is called the characteristic frequency. Supra-

threshold sound stimulation activates neurons with tuning curves that lie over a broader

frequency range. Thus, it is not implausible that we see changes in tonotopic maps at a

broader range of frequencies following the 70 dB SPL, 2-kHz-centered narrowband

sound stimulation.

Page 113: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

102

Figure 6.1: Schematic of chinchilla frequency tuning curves spanning a range of

frequencies. Note how tuning curve shape varies with frequency. The red oval

indicates the 70 dB SPL, 2-kHz-centered narrowband sound exposure stimulus. Cross-

hatching indicates neural tuning curves that would be activated by the sound-exposure

stimulus, for the given neural tuning curve thresholds shown.

Results from electrophysiological experiments suggest that there is an increase in

the neural representation of frequencies beginning approximately 1 kHz below the 2 ±

0.25 kHz sound-exposure stimulus, spanning 0.1 – 1 kHz. While it has been previously

reported that early post-natal exposure to an enhanced acoustic environment increases

cortical representation of the octave band containing the exposure frequencies (Stanton

and Harrison 1996), these experiments were carried out in an altricious species, compared

with the precocious chinchilla used in the present experiments. A critical examination of

the literature suggests that increased cortical and sub-cortical (inferior colliculus) areas

dedicated to frequencies below the exposure frequency is not uncommon following post-

natal rearing in an enhanced acoustic environment.

Page 114: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

103

In an early report by Clopton and Winfield (1976), rats were exposed for the first

4 months of life to a 1-s frequency sweep (“up”: 6 – 9 kHz, “down”: 9 – 6 kHz; 65 dB

SPL) followed by a 1-s noise burst at 40 dB SPL. This pattern was repeated for an

average of 5 hours/day. Units in inferior colliculus were sampled. Data from Figures 3

and 5 of their paper (Clopton and Winfield 1976) are shown re-plotted in Figure 6.2. The

greatest number of responses occurred between 4 and 5 kHz; that is, about 1 kHz below

the lowest frequency of the exposure sweep.

Characteristic Frequency (kHz)

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16

Re

sp

on

se

s

0

2

4

6

8

10

12

14

Up-exposed

Down-exposed

Figure 6.2: Data from Figures 3 & 5 of Clopton & Winfield (1976) re-plotted. Subjects

were exposed for the first 4 months of life to patterned sound: a 1-s upward (or

downward) tone sweep from 6 – 9 kHz (or 9 - 6 kHz) at 65 dB SPL, followed by a 1-s

40-dB-SPL noise burst, repeated for an average of 5 hours/day. The majority of responses

occurred at characteristic frequencies between 4- and 5-kHz, 1-kHz below the lower limit

of the exposure tones. The black horizontal line indicates frequencies in the exposure

sweep.

Page 115: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

104

More recently, close inspection of Fig. 2 in Oliver et al. (2011), shows increased

representation of 13 kHz in adult rat IC (recordings made at P212) following 60-70 dB

exposure to a 14-kHz pure tone (25 ms, every 250 ms) from P9 – P28 (Figure 6.3).

Control

Augmented Acoustic Environment

Figure 6.3: Increased neural representation of frequencies about 1-kHz below the tone-

exposure frequency of 14 kHz (indicated by dashed lines), recorded from various

electrode tracks (TR) in a control (162ctrl) and sound-exposed (AAE83) subject. Adapted

from Oliver et al., 2011.

In a study of primary auditory cortex by Merzenich and colleagues (Nakahara et

al., 2004), rats were exposed to a sequence of 30-ms 65-dB-SPL tone pips from P9 to

P30. The first sequence consisted of 2.8-, 5.6-, 4.4-kHz pips that began at 0, 150, and 300

ms, respectively; the second sequence: 15-, 21-, and 30-kHz pips at 500, 650, and 800

ms, respectively. The greatest percent change in cortical representation of adult rats was

at 1.7 kHz, about an octave below the lowest tone stimulus presented, with a 140%

increase in cortical area dedicated to these frequencies (bin width, 0.6 octaves) compared

(14 kHz)

14 kHz

14 kHz

Fre

quen

cy (

kH

z)

Fre

quen

cy (

kH

z)

Relative penetration depth (mm) Relative penetration depth (mm)

16

14

12

0.0 0.2 0.4 0.0 0.2 0.4

162ctrlTR1

162ctrlTR2

162ctrlTR3

162ctrlTR5

AAE83TR1

AAE83TR2

AAE83TR3

13 kHz

13 kHz

13 kHz

16

14

12

16

14

12

16

14

12

16

14

12

16

14

12

16

14

12

Page 116: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

105

with controls. These studies provide evidence that developmental changes in tonotopic

maps following neonatal exposure to tonal stimulation are not limited to the exposure

frequency. Changes at low-frequency “edges” of the sound stimulus have been reported.

6.2 Global neuronal expression patterns: c-fos immunolabeling

6.2.1 Technique development

We have developed a reliable method to label c-fos-active cells in auditory

midbrain of chinchilla. Our protocol is based on methods similar to those developed for

other animal species (Brown and Liu 1995; Harrison and Negandhi 2012; Lu et al.,

2009). However, these studies focus on altricious laboratory subjects (e.g., mouse, rat),

who undergo hearing onset after birth (P12-P14 in rat; Geal-Dor et al., 1993), whereas

the chinchilla is precocious: cochlear function begins in utero, similar to humans

(Querleu et al., 1988). Moreover, our protocol uses only commercially-available reagents,

thus it is accessible to the wider research community.

To use c-fos labeling to reveal functional organization in the auditory midbrain:

the subject must be maintained in a (near) silent environment for many hours prior to

hearing the acoustic stimulus of interest; the subject has to hear the stimulus for an

adequate period of time; and the brain must be fixed at within a time window

corresponding to the dynamics of c-fos expression. Our protocol has achieved this. We

can histologically image tonotopic bands present following pure-tone sound stimulation.

The location of these bands is consistent with the known sound-frequency organization of

auditory midbrain. The presence of bands agrees with reports in the literature from other

Page 117: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

106

animal species measured both with c-fos immunolabeling (e.g., Brown and Liu 1995;

Friauf 1992) and 2-deoxyglucose measurements (e.g., Ehret and Romand 1994). To our

knowledge, the correlation of these bands with electrophysiological recordings, and the

potentially inhibitory sidebands observed adjacent to these bands are new findings.

Taken together, these results suggest that the c-fos labeling technique can be used to

study tonotopic organization in chinchilla.

Although the relationship between neuronal activation and immediate early gene

induction is not fully understood, part of the signal transduction pathway involves the

influx of Ca2+ either via activation of NMDA-type glutamate receptors or via L-type

voltage-gated Ca2+ channels (Platenik et al., 2000; Sheng and Greenberg 1990). The c-fos

protein is translated from c-fos mRNA in the cytoplasm, then moves into the cell nucleus.

It is at the level of the cell nucleus that the c-fos protein is labeled in the present thesis

(Curran et al., 1984; Roux et al., 1990). In the cell nucleus, the c-fos protein dimerizes

with other transcription factors (members of the Jun family) to form part of the activator

protein-1 (AP-1) transcription factor (Curran and Franza 1988). AP-1 upregulates

transcription of a wide variety of genes involved in cell proliferation and differentiation

(Angel & Karin 1991), tumorigenesis, and apoptosis (Preston et al. 1996).

In addition and of some interest, we have observed regions of decreased cell

labeling adjacent to the bands of labeled neurons, which has not been previously reported

using this technique. It is plausible that this decrease represents an inhibitory region.

Lateral (or sideband) inhibition is a rather ubiquitous property of auditory neurons and

has been studied at the level of IC by a number of authors (Alkhatib et al., 2006; Ehret

and Merzenich 1988; Palombi and Caspary 1996a). Furthermore, iontophoretic

Page 118: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

107

application of bicuculline and strychnine -- GABAA and glycine receptor antagonists,

respectively -- increases firing rate and tuning curve width of neurons in IC (LeBeau et

al., 2001; Mayko et al., 2012; Xie et al., 2005). Anatomically, neurons in the central

nucleus of inferior colliculus are arranged in iso-frequency laminae that are orthogonal to

the tonotopic axis (Morest and Oliver 1984; Oliver and Morest 1984; Schreiner and

Langner 1997). Malmierca et al. (1993) describe two types of neurons observed in the IC

of adult rat: flat neurons (called disc-shaped neurons in other species) have dendritic

arbors that are thinner and denser, and form laminae; less flat neurons (stellate cells, in

other species) have thicker arbors, are less dense, and are found to populate the inter-

laminar compartments. Regions of inhibitory activity may occur in these interlaminar

compartments: inhibitory connections between laminae have been reported in the bat

(Pollak and Park 1993) and cat (Palombi and Caspary 1996b). While it is reasonable to

posit that the regions of decreased cell labeling adjacent to the bands of fos-labeled cells

correspond to inhibitory regions, further studies are needed to test this hypothesis.

Various studies have examined how brief periods of sensory stimulation alter c-

fos expression patterns. In somatosensory cortex, vibrissae stimulation has been shown to

increase c-fos expression in rat barrel cortex. Following 15-20 minutes of manual

whisker stimulation in anaesthetized (Mack & Mack 1992) and awake subjects (trained to

sit atop a copper cylinder; Filipkowski et al. 2000), c-fos expression levels in barrel

cortex were increased relative to unstimulated controls. For more “natural” whisker

stimulation, in a second experimental paradigm, subjects were placed in a new wire cage

(where they exhibited increased sniffing and whisking behaviour). Again, significant c-

fos expression was reported in barrel cortex in these subjects compared with controls

Page 119: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

108

(Filipkowski et al. 2000). In visual cortex, experiments to study the effects of sensory

stimulation on IEG expression often involve delivering visual input following a period of

dark-rearing (review: Kaczmarek & Chaudhuri 1997). C-fos expression levels in visual

cortex were increased in dark-reared kittens (birth to 5 weeks, then 1 hr of light

stimulation: Rosen et al. 1992; birth to 30 days, then 0.5, 1, 2, 4, or 6 hrs of light

stimulation: Beaver et al. 1993). Maximum c-fos induction occurred following the 2-hr

period of light stimulation (Beaver et al. 1993). In the auditory system, brief (1-2 hr)

pure-tone sound stimulation induces tonotopic bands in a number of animal models

(mouse: Brown and Liu 1995; rat: Lu et al., 2009), and is the stimulation duration used in

the present thesis to elicit bands of c-fos-labeled neurons.

We note that the spatial extent of the region of c-fos labeled cells is more than

would be expected if a near threshold tonal stimulus was used for c-fos activation. Our

stimulation tone to induce bands of fos-labeled cells was presented suprathrehold (at 60

dB SPL), and thus activated a wider neuron array (Figure 6.1). It is also possible that the

c-fos labeled band width is determined by the laminar neuro-architecture of IC central

nucleus as described by Malmierca et al. (1993) and Winer and Schreiner (2005). This

laminar neuronal organization in IC translates functionally into "frequency band

laminae", and may relate to observations of a discontinuous, or stepwise increment in

tonotopic organization (Malmierca et al., 2008; Schreiner and Langner 1997), which has

been reported as ~0.3 octaves per frequency step, independent of frequency (rat IC: 0.29

octaves per frequency step, Malmierca et al. 2008; cat IC: 0.28 octaves per frequency

step, Schreiner & Langner 1997).

Page 120: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

109

Immunohistochemical techniques complement electrophysiological studies. Much

progress has been made in understanding neural function throughout the brain with multi-

electrode, multi-site recordings in both anaesthetized and awake subjects. While

electrophysiological studies have high temporal resolution, they are invasive, and have

low spatial coverage. Previous histological activity markers that have been employed

(e.g., [C14]-2-deoxyglucose, and the mitochrondrial enzyme cytochrome oxidase) have

supplemented results from electrophysiological methodologies by permitting

visualization of global neuronal activation patterns. These methods have high spatial

resolution, but low temporal resolution. C-fos immunohistochemistry also allows

mapping of global neural activation patterns, with the additional benefit of single-cell-

level resolution. Additionally, during stimulation and subsequent induction of immediate

early genes, subjects are awake, with minimal interruption to their “normal” behavior. In

the present study, we have employed these complementary techniques; results observed

from each technique are not incompatible.

6.2.2 c-fos expression patterns following sound-exposure

Subjects who heard the sound exposure stimulus for 4 weeks, then received no

additional sound stimulation did not exhibit c-fos expression patterns that differed from

age-matched controls (Figure 5.4C & D; Figure 5.6). These results imply that sound-

exposure does not change basal levels of neural activity in inferior colliculus. This

includes neurons that are spontaneously active, as well as those that are auto-evoked

(e.g., from the subject’s movement within the cage, breathing rhythms). These results are

in accordance with qualitative electrophysiological results, seen in the tuning curves of

Page 121: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

110

Figure 4.6, which also show that off-frequency levels of neural activity were not different

between groups over a broad frequency range.

Sound-exposed subjects who subsequently heard the 90-min, 2-kHz pulsed tone

exhibited increased c-fos expression levels, both near the 2-kHz region of interest and

over a broad range of inferior colliculus (Figure 5.4A & B; Figure 5.5). It is important to

note here the difference in timing between the two main experimental techniques used in

this thesis (Figure 6.4). Electrophysiological recordings were made in IC neurons of

sound-exposed subjects in response to 50-ms tone pips (Figure 6.4A), while sound-

exposed subjects in this set of c-fos expression experiments essentially underwent 2

experimental “treatments”: the 4-week sound exposure, then the 90-min, 2-kHz pulsed

tone exposure (to induce bands of fos-labeled neurons, Figure 6.4B). Taken together with

the results of the previous paragraph, these data suggest that the sound-exposure stimulus

alters the way in which neurons respond to a subsequent, prolonged presentation of a

sound stimulus in the frequency range of the sound exposure stimulus (90 min pulsed 2-

kHz tone; shown schematically in Figure 6.4B).

Page 122: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

111

Figure 6.4: Schematic emphasizing the difference in timing between (A) micro-

electrode electrophysiological recordings, and (B) c-fos immunohistochemical

experiments for sound-exposed subjects who subsequently heard the 90-min, 2-kHz

pulsed pure-tone stimulus.

While the mechanism by which extracellular stimuli result in induction of the c-

fos immediate early gene is not clear, part of the signal transduction pathway is known to

include activation of NMDA receptors (Platenik et al., 2000; Sheng and Greenberg

1990). During development, there are changes in glutamatergic transmission. “Silent

synapses” (those that contain only NMDA receptors) have been proposed to predominate

in early development (Durand et al., 1996; Wu et al., 1996). Sensory experience and

NMDAR activation converts these synapses to functional ones via NMDA-receptor-

mediated insertion of AMPA receptors (Isaac et al., 1997; Quinlan et al., 1999).

Additionally, NMDAR sub-unit composition switches: NMDARs composed of NR1 and

NR2B sub-units dominate in sensory cortices; during development, NR1- and NR2A-

containing sub-units emerge (Quinlan et al., 1999; Sheng et al., 1994). It has been

A

B

Sound-Exposure

Sound-Exposure Silence

ABR, then

SEM or

Micro-electrode recordings (50-ms tone pips)

Birth

Birth P28 P27

P28

ABR

90 min, 2-kHz sound stimulation

c-fos immunohistochemistry

Page 123: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

112

suggested that NMDAR sub-unit composition influences the degree of synaptic plasticity

during both developmental and adult periods, since Ca2+ permeability is determined by

subunit composition (review: Lau et al., 2009). Moreover, synaptic transmission can be

maintained in an immature state by delaying sensory experience (e.g., dark-rearing in the

visual system: Quinlan et al., 1999).

In a study by Hogsden and Dringenberg (2009), the authors reliably induced long-

term potentiation (LTP) in rat primary auditory cortex by theta-burst stimulation of the

medial geniculate nucleus (of thalamus). During early post-natal life, a second set of

subjects were reared in continuous white noise (from P5 to P50-P60; 70-80 dB SPL),

which effectively masked patterned auditory input. LTP was enhanced in these white-

noise-reared subjects compared to age-matched controls reared in an unaltered sound

environment. The enhanced LTP was abolished following application of NR2B-subunit

antagonists (Ro 25-6981 or ifenprodil). The authors suggest that post-natal white-noise

exposure maintains higher, juvenile-like levels of synaptic plasticity. It is plausible that

the sound-exposure stimulus used in the present thesis preserved regions of the IC in an

immature state (greater proportion of NR2B sub-unit composition), and that subsequent

stimulation – of sufficient duration – induced an enhanced neural response, similar to that

reported by Hogsden and Dringenberg (2009), seen here as an increase in the number of

c-fos labeled neurons (Figure 5.4A & B; Figure 5.5).

6.3 Methodological considerations

The results reported herein should be viewed in light of various methodological

considerations. First, electrophysiological results were obtained from anaesthetized

subjects. There are several advantages to this experimental preparation: factors such as

Page 124: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

113

attention and comfort level that would be present in awake subjects are eliminated; stable

recordings can be made for longer periods of time in anaesthetized subjects; and with the

chosen anaesthetic regime (ketamine/xylazine), subjects respire spontaneously,

eliminating the need for artificial respiration, which facilitates the brainstem stability

needed for micro-electrode recordings. Considerations include that the consistent

administration of intraperitoneal anaesthetic may be subject to error. Care must be taken

when comparing results from awake subjects to those from anaesthetized subjects, and

also when comparing results from studies employing differing anaesthetic regimes. For

example, in a study of response properties of IC neurons in the guinea pig, Syka and

colleagues (Astl et al., 1996) report that spontaneous activity was increased in ketamine-

anaesthetized subjects compared with urethane- or pentobarbital-anaesthetized subjects.

Thresholds of units with a sustained response were lower in ketamine- and urethane-

anaesthetized subjects than units with an onset response. Frequency tuning was not

different between the anaesthetic types used (Astl et al., 1996).

Second, electrophysiological data were recorded, for the most part, from multi-

units. We estimate to have recorded from approximately 1-5 units per electrode at each

electrode depth. Recall that characteristic frequency increases with increasing electrode

depth in the inferior colliculus. Given this tonotopic arrangement, and that electrode

tracks were oriented orthogonal to the tonotopic axis, it is probable that at a given

electrode depth, the few neurons from which we recorded would have similar

characteristic frequencies. Also, neural responses from collicular neurons are quite

robust, extending well beyond the noise floor (Figure 3.7), making spike discrimination

relatively straight-forward with a simple voltage-thresholding technique. This differs

Page 125: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

114

from neural discrimination at the level of the cortex, for example, where neural responses

are not as easily discernible from the noise floor, and more complex analysis techniques

are needed to discern neural activity from electrode noise (such as statistical comparisons

between noise floor and electrode voltages, e.g., Brown and Harrison 2009).

Regarding the c-fos immunohistochemical technique, we re-iterate its low

temporal resolution: c-fos protein synthesis peaks 30-90 minutes after stimulation, and

has a half-life of a few hours (Morgan and Curran 1991; Herdegen and Leah 1998). Thus,

the technique is not ideally suited for short-lasting behavioural states, nor when

stimulation lasts longer than several hours. In the present set of experiments, we studied

changes in neural activation patterns to a 90-min tone stimulus before and after the 4-

week sound-exposure period, thus working within the time frame of highest c-fos protein

expression. C-fos studies were complemented with micro-electrode recordings to

measure changes in neural response patterns to short-duration (50-ms) tones, which

would not have been possible with the low temporal resolution of the c-fos technique.

It is important to note that c-fos protein expression is not initiated by changes in

cell energy metabolism unlike, for example, 2-deoxyglucose uptake. We mention this

because it is sometimes debated as to whether inhibitory processes use less or equivalent

energy to neural excitation. If c-fos labeling was dependent on ATP use, we would need

to consider this issue. However, it is accepted that the initiation of c-fos expression is

triggered by calcium channel modulation directly associated with cell excitation

(depolarization; reviewed in Finkbeiner & Greenberg 1998) and not linked to energy

metabolism. C-fos immunohistochemistry does not label neurons whose activity is

inhibited.

Page 126: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

115

6.4 Conclusions, Future Work

The data presented in this thesis support the hypothesis that changes in sound

stimulation patterns at the periphery (the ear) affect change in central sensory

representations. We provided evidence that the sound exposure stimulus does not alter

cochlear function or structure: in the frequency range surrounding the sound-exposure

frequencies, we observed similar ABR thresholds, hair cell morphology, and neural

thresholds between control and sound-exposed subjects. Prolonged (4-week), passive

exposure to the sound exposure stimulus was shown to decrease neural representation of

2- & 8-kHz octave bands, and increase neural representation of low-frequency regions

(< 1 kHz) in auditory midbrain.

Compared with age-matched controls, we observed a significant increase in the

number of c-fos-labeled neurons both in the 2-kHz region (corresponding to the sound-

exposure frequency) and over a broad frequency range in inferior colliculus of sound-

exposed subjects who subsequently heard the 90 min, 2-kHz pulsed pure-tone. There

were no differences in c-fos expression in sound-exposed subjects who heard no

additional sound, suggesting that sound-exposure does not alter basal levels of neural

activity. These data support the hypothesis that persistent exposure to an abnormal sound

environment -- during an early post-natal period -- can alter tonotopic organization in

auditory midbrain. More broadly, this study contributes to a growing body of literature

that suggests that sub-cortical levels of the ascending auditory pathway reorganize

following changes in peripheral input during development.

Several additional research questions arise from this work. Are the changes in

sound frequency representation following sound exposure reported in this thesis long-

Page 127: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

116

lasting? That is, would they be present several weeks or months following the end of the

sound-exposure period? Answers to these questions would give us insight as to whether

sound exposure maintains the auditory system in an immature state, effectively extending

the critical period, as reported in Chang & Merzenich (2003). Can returning the subjects

to a “normal” sound environment reverse the changes reported herein, thus restoring

normal sound processing, as in Zhou et al. (2011)? Also, our experiments were carried

out in developing subjects. It remains to be seen whether passive exposure to an

enhanced sound environment in a precocious adult subject suffices to alter neural

responsiveness. We have seen some evidence for this in mature altricious subjects

(Pienkowski and Eggermont 2010). Further experiments are needed to explore these

research questions.

Page 128: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

117

References

Aitkin, L., Tran, L., Syka, J., 1994. The responses of neurons in subdivisions of the

inferior colliculus of cats to tonal, noise and vocal stimuli. Exp Brain Res. 98, 53-64.

Aitkin, L.M., Anderson, D.J., Brugge, J.F., 1970. Tonotopic organization and discharge

characteristics of single neurons in nuclei of the lateral lemniscus of the cat. J

Neurophysiol. 33, 421-440.

Alkhatib, A., Biebel, U.W., Smolders, J.W., 2006. Inhibitory and excitatory response

areas of neurons in the central nucleus of the inferior colliculus in unanesthetized

chinchillas. Exp Brain Res. 174, 124-143.

Angel, P., Karin, M., 1991. The role of jun, fos and the AP-1 complex in cell-

proliferation and transformation. Biochim Biophys Acta. 1072, 129-157.

Astl, J., Popelář, J., Kvašňák, E., Syka, J., 1996. Comparison of response properties of

neurons in the inferior colliculus of guinea pigs under different anesthetics. Audiology.

35, 335-345.

Bao, S., Chan, V.T., Merzenich, M.M., 2001. Cortical remodelling induced by activity of

ventral tegmental dopamine neurons. Nature. 412, 79-83.

Beaver, C.J., Mitchell, D.E., Robertson, H.A., 1993. Immunohistochemical study of the

pattern of rapid expression of C-fos protein in the visual cortex of dark-reared kittens

following initial exposure to light. J Comp Neurol. 333, 469-484.

Bi, G.Q., Poo, M.M., 1998. Synaptic modifications in cultured hippocampal neurons:

Dependence on spike timing, synaptic strength, and postsynaptic cell type. J Neurosci.

18, 10464-10472.

Birnholz, J.C., Benacerraf, B.R., 1983. The development of human fetal hearing. Science.

222, 516-518.

Page 129: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

118

Bliss, T.V., Lomo, T., 1973. Long-lasting potentiation of synaptic transmission in the

dentate area of the anaesthetized rabbit following stimulation of the perforant path. J

Physiol. 232, 331-356.

Brainard, M.S., Knudsen, E.I., 1993. Experience-dependent plasticity in the inferior

colliculus: A site for visual calibration of the neural representation of auditory space in

the barn owl. J Neurosci. 13, 4589-4608.

Bramham, C.R., Worley, P.F., Moore, M.J., Guzowski, J.F., 2008. The immediate early

gene arc/arg3.1: Regulation, mechanisms, and function. J Neurosci. 28, 11760-11767.

Brown, M.C., Liu, T.S., 1995. Fos-like immunoreactivity in central auditory neurons of

the mouse. J Comp Neurol. 357, 85-97.

Brown, T.A., Harrison, R.V., 2011. Neuronal responses in chinchilla auditory cortex after

postnatal exposure to frequency-modulated tones. Hear Res. 275, 8-16.

Brown, T.A., Harrison, R.V., 2010. Postnatal development of neuronal responses to

frequency-modulated tones in chinchilla auditory cortex. Brain Res. 1309, 29-39.

Brown, T.A., Harrison, R.V., 2009. Responses of neurons in chinchilla auditory cortex to

frequency-modulated tones. J Neurophysiol. 101, 2017-2029.

Calford, M.B., Aitkin, L.M., 1983. Ascending projections to the medial geniculate body

of the cat: Evidence for multiple, parallel auditory pathways through thalamus. J

Neurosci. 3, 2365-2380.

Caporale, N., Dan, Y., 2008. Spike timing-dependent plasticity: A hebbian learning rule.

Annu Rev Neurosci. 31, 25-46.

Chang, E.F., Bao, S., Imaizumi, K., Schreiner, C.E., Merzenich, M.M., 2005.

Development of spectral and temporal response selectivity in the auditory cortex. Proc

Natl Acad Sci U S A. 102, 16460-16465.

Chang, E.F., Merzenich, M.M., 2003. Environmental noise retards auditory cortical

development. Science. 300, 498-502.

Clopton, B.M., Winfield, J.A., 1976. Effect of early exposure to patterned sound on unit

activity in rat inferior colliculus. J Neurophysiol. 39, 1081-1089.

Cochran, B.H., Zullo, J., Verma, I.M., Stiles, C.D., 1984. Expression of the c-fos gene

and of an fos-related gene is stimulated by platelet-derived growth factor. Science. 226,

1080-1082.

Curran, T., Franza, B.R.,Jr, 1988. Fos and jun: The AP-1 connection. Cell. 55, 395-397.

Page 130: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

119

Curran, T., Miller, A.D., Zokas, L., Verma, I.M., 1984. Viral and cellular fos proteins: A

comparative analysis. Cell. 36, 259-268.

D'Alessandro, L.M., Harrison, R.V., 2014. Excitatory and inhibitory tonotopic bands in

chinchilla inferior colliculus revealed by c-fos immuno-labeling. Hear Res. 316, 122-128.

de Villers-Sidani, E., Simpson, K.L., Lu, Y.F., Lin, R.C., Merzenich, M.M., 2008.

Manipulating critical period closure across different sectors of the primary auditory

cortex. Nat Neurosci. 11, 957-965.

de Villers-Sidani, E., Chang, E.F., Bao, S., Merzenich, M.M., 2007. Critical period

window for spectral tuning defined in the primary auditory cortex (A1) in the rat. J

Neurosci. 27, 180-189.

DeBello, W.M., Feldman, D.E., Knudsen, E.I., 2001. Adaptive axonal remodeling in the

midbrain auditory space map. J Neurosci. 21, 3161-3174.

Durand, G.M., Kovalchuk, Y., Konnerth, A., 1996. Long-term potentiation and

functional synapse induction in developing hippocampus. Nature. 381, 71-75.

Ehret, G., Romand, R., 1994. Development of tonotopy in the inferior colliculus II: 2-DG

measurements in the kitten. Eur J Neurosci. 6, 1589-1595.

Ehret, G., Merzenich, M.M., 1988. Complex sound analysis (frequency resolution,

filtering and spectral integration) by single units of the inferior colliculus of the cat. Brain

Res. 472, 139-163.

Evans, E.F., 1982. Funtional anatomy of the auditory system. In The Senses. H.B.

Barlow, J.D. Mollon, eds. pp. 251-306.

Evans, E.F., 1972. The frequency response and other properties of single fibres in the

guinea-pig cochlear nerve. J Physiol. 226, 263-287.

Fettiplace, R., Hackney, C.M., 2006. The sensory and motor roles of auditory hair cells.

Nat Rev Neurosci. 7, 19-29.

Filipkowski, R.K., Rydz, M., Berdel, B., Morys, J., Kaczmarek, L., 2000. Tactile

experience induces c-fos expression in rat barrel cortex. Learn Mem. 7, 116-122.

Finkbeiner, S., Greenberg, M.E., 1998. Ca2+ channel-regulated neuronal gene expression.

J Neurobiol. 37, 171-189.

Franklin, K.B.J., Paxinos, G., 2008, The Mouse Brain in Stereotaxic Coordinates.

Academic Press, New York, USA.

Page 131: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

120

Friauf, E., 1992. Tonotopic order in the adult and developing auditory system of the rat as

shown by c-fos immunocytochemistry. Eur J Neurosci. 4, 798-812.

Geal-Dor, M., Freeman, S., Li, G., Sohmer, H., 1993. Development of hearing in

neonatal rats: Air and bone conducted ABR thresholds. Hear Res. 69, 236-242.

Greenberg, M.E., Hermanowski, A.L., Ziff, E.B., 1986. Effect of protein synthesis

inhibitors on growth factor activation of c-fos, c-myc, and actin gene transcription. Mol

Cell Biol. 6, 1050-1057.

Harel, N., Mori, N., Sawada, S., Mount, R.J., Harrison, R.V., 2000. Three distinct

auditory areas of cortex (AI, AII, and AAF) defined by optical imaging of intrinsic

signals. Neuroimage. 11, 302-312.

Harrison, A.L., Negandhi, J., Allemang, C., D'Alessandro, L.M., Harrison, R.V., 2013.

Acoustic cross-over between the ear in mice (Mus musculus) determined using a novel

ABR based bio-assay. Canadian Acoustics. 41, 5-10.

Harrison, R.V., Negandhi, J., 2012. Resting neural activity patterns in auditory brainstem

and midbrain in conductive hearing loss. Acta Otolaryngol. 132, 409-414.

Harrison, R.V., 2006. Development, maintenance and plasticity of tonotopic projections

from cochlea to auditory cortex. In Reprogramming the Cerebral Cortex: Plasticity

Following Central and Peripheral Lesions. S.G. Lomber, J.J. Eggermont, eds. Oxford

University Press Inc., New York, USA, pp. 159-180.

Harrison, R.V., 2001. Age-related tonotopic map plasticity in the central auditory

pathways. Scand Audiol Suppl. (53), 8-14.

Harrison, R.V., Ibrahim, D., Mount, R.J., 1998. Plasticity of tonotopic maps in auditory

midbrain following partial cochlear damage in the developing chinchilla. Exp Brain Res.

123, 449-460.

Harrison, R.V., Cullen, J.R., Takeno, S., Mount, R.J., 1996. The neonatal chinchilla

cochlea: Morphological and functional study. Scanning Microsc. 10, 889-894.

Harrison, R.V., Stanton, S.G., Mount, R.J., 1995. Effects of chronic cochlear damage on

threshold and frequency tuning of neurons in AI auditory cortex. Acta Otolaryngol Suppl.

519, 30-35.

Hebb, D.O., 1949, The Organization of Behaviour. Wiley & Sons, New York.

Heffner, R.S., Heffner, H.E., 1991. Behavioral hearing range of the chinchilla. Hear Res.

52, 13-16.

Page 132: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

121

Herdegen, T., Leah, J.D., 1998. Inducible and constitutive transcription factors in the

mammalian nervous system: Control of gene expression by jun, fos and krox, and

CREB/ATF proteins. Brain Res Brain Res Rev. 28, 370-490.

Herrera, D.G., Robertson, H.A., 1996. Activation of c-fos in the brain. Prog Neurobiol.

50, 83-107.

Hogsden, J.L., Dringenberg, H.C., 2009. NR2B subunit-dependent long-term potentiation

enhancement in the rat cortical auditory system in vivo following masking of patterned

auditory input by white noise exposure during early postnatal life. Eur J Neurosci. 30,

376-384.

Hosomi, H., Hirai, H., Okada, Y., Amatsu, M., 1995. Long-term potentiation of

neurotransmission in the inferior colliculus of the rat. Neurosci Lett. 195, 175-178.

Huang, C.L., Winer, J.A., 2000. Auditory thalamocortical projections in the cat: Laminar

and areal patterns of input. J Comp Neurol. 427, 302-331.

Hughes, P., Dragunow, M., 1995. Induction of immediate-early genes and the control of

neurotransmitter-regulated gene expression within the nervous system. Pharmacol Rev.

47, 133-178.

Imig, T.J., Morel, A., 1984. Topographic and cytoarchitectonic organization of thalamic

neurons related to their targets in low-, middle-, and high-frequency representations in cat

auditory cortex. J Comp Neurol. 227, 511-539.

Isaac, J.T., Crair, M.C., Nicoll, R.A., Malenka, R.C., 1997. Silent synapses during

development of thalamocortical inputs. Neuron. 18, 269-280.

Jeffress, L.A., 1948. A place theory of sound localization. J Comp Physiol Psychol. 41,

35-39.

Jenkins, W.M., Masterton, R.B., 1982. Sound localization: Effects of unilateral lesions in

central auditory system. J Neurophysiol. 47, 987-1016.

Kaczmarek, L., Chaudhuri, A., 1997. Sensory regulation of immediate-early gene

expression in mammalian visual cortex: Implications for functional mapping and neural

plasticity. Brain Res Brain Res Rev. 23, 237-256.

Kakigi, A., Hirakawa, H., Harel, N., Mount, R.J., Harrison, R.V., 2000. Tonotopic

mapping in auditory cortex of the adult chinchilla with amikacin-induced cochlear

lesions. Audiology. 39, 153-160.

Kaltenbach, J.A., Meleca, R.J., Falzarano, P.R., 1996. Alterations in the tonotopic map of

the cochlear nucleus following cochlear damage. In Auditory System Plasticity and

Page 133: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

122

Regeneration. R.J. Salvi, D. Henderson, F. Fiorino, V. Colletti, eds. Thieme Medical

Publishers, Inc., New York, USA, pp. 317-332.

Katz, J., Lezynski, J., 2002. Clinical masking. In Handbook of Clinical Audiology. J.

Katz, ed. Lippincott Williams & Wilkins, Baltimore, USA, pp. 124-141.

Kauer, J.A., Malenka, R.C., 2007. Synaptic plasticity and addiction. Nat Rev Neurosci. 8,

844-858.

Kilgard, M.P., Merzenich, M.M., 1998. Cortical map reorganization enabled by nucleus

basalis activity. Science. 279, 1714-1718.

Knapska, E., Kaczmarek, L., 2004. A gene for neuronal plasticity in the mammalian

brain: Zif268/egr-1/NGFI-A/krox-24/TIS8/ZENK? Prog Neurobiol. 74, 183-211.

Knudsen, E.I., 2004. Sensitive periods in the development of the brain and behavior. J

Cogn Neurosci. 16, 1412-1425.

Kujawa, S.G., Liberman, M.C., 2009. Adding insult to injury: Cochlear nerve

degeneration after "temporary" noise-induced hearing loss. J Neurosci. 29, 14077-14085.

Lau, C.G., Takeuchi, K., Rodenas-Ruano, A., Takayasu, Y., Murphy, J., Bennett, M.V.,

Zukin, R.S., 2009. Regulation of NMDA receptor Ca2+ signalling and synaptic plasticity.

Biochem Soc Trans. 37, 1369-1374.

LeBeau, F.E., Malmierca, M.S., Rees, A., 2001. Iontophoresis in vivo demonstrates a key

role for GABA(A) and glycinergic inhibition in shaping frequency response areas in the

inferior colliculus of guinea pig. J Neurosci. 21, 7303-7312.

LeDoux, J.E., Ruggiero, D.A., Reis, D.J., 1985. Projections to the subcortical forebrain

from anatomically defined regions of the medial geniculate body in the rat. J Comp

Neurol. 242, 182-213.

Liberman, M.C., Dodds, L.W., 1984. Single-neuron labeling and chronic cochlear

pathology. III. stereocilia damage and alterations of threshold tuning curves. Hear Res.

16, 55-74.

Loebrich, S., Nedivi, E., 2009. The function of activity-regulated genes in the nervous

system. Physiol Rev. 89, 1079-1103.

Lomber, S.G., Eggermont, J.J., 2006, Reprogramming the Cerebral Cortex: Plasticity

Following Central and Preipheral Lesions. Oxford University Press, New York.

Lu, H.P., Chen, S.T., Poon, P.W., 2009. Nuclear size of c-fos expression at the auditory

brainstem is related to the time-varying nature of the acoustic stimuli. Neurosci Lett. 451,

139-143.

Page 134: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

123

Lyford, G.L., Yamagata, K., Kaufmann, W.E., Barnes, C.A., Sanders, L.K., Copeland,

N.G., Gilbert, D.J., Jenkins, N.A., Lanahan, A.A., Worley, P.F., 1995. Arc, a growth

factor and activity-regulated gene, encodes a novel cytoskeleton-associated protein that is

enriched in neuronal dendrites. Neuron. 14, 433-445.

Ma, C.L., Kelly, J.B., Wu, S.H., 2002. AMPA and NMDA receptors mediate synaptic

excitation in the rat's inferior colliculus. Hear Res. 168, 25-34.

Mack, K.J., Mack, P.A., 1992. Induction of transcription factors in somatosensory cortex

after tactile stimulation. Brain Res Mol Brain Res. 12, 141-147.

Malmierca, M.S., Izquierdo, M.A., Cristaudo, S., Hernandez, O., Perez-Gonzalez, D.,

Covey, E., Oliver, D.L., 2008. A discontinuous tonotopic organization in the inferior

colliculus of the rat. J Neurosci. 28, 4767-4776.

Malmierca, M.S., Hernandez, O., Rees, A., 2005. Intercollicular commissural projections

modulate neuronal responses in the inferior colliculus. Eur J Neurosci. 21, 2701-2710.

Malmierca, M.S., Hernandez, O., Falconi, A., Lopez-Poveda, E.A., Merchan, M., Rees,

A., 2003. The commissure of the inferior colliculus shapes frequency response areas in

rat: An in vivo study using reversible blockade with microinjection of kynurenic acid.

Exp Brain Res. 153, 522-529.

Malmierca, M.S., Blackstad, T.W., Osen, K.K., Karagulle, T., Molowny, R.L., 1993. The

central nucleus of the inferior colliculus in rat: A golgi and computer reconstruction study

of neuronal and laminar structure. J Comp Neurol. 333, 1-27.

Mayko, Z.M., Roberts, P.D., Portfors, C.V., 2012. Inhibition shapes selectivity to

vocalizations in the inferior colliculus of awake mice. Front Neural Circuits. 6, 73.

Mello, C.V., Pinaud, R., 2006. Immediate early gene regulation in the auditory system. In

Immediate Early Genes in Sensory Processing, Cognitive Performance and Neurological

Disorders. R. Pinaud, L.A. Tremere, eds. Springer Verlag, New York, USA, pp. 35-56.

Merzenich, M.M., Reid, M.D., 1974. Representation of the cochlea within the inferior

colliculus of the cat. Brain Res. 77, 397-415.

Miller, J.D., 1970. Audibility curve of the chinchilla. J Acoust Soc Am. 48, 513-523.

Miyakawa, A., Gibboni, R., Bao, S., 2013. Repeated exposure to a tone transiently alters

spectral tuning bandwidth of neurons in the central nucleus of inferior colliculus in

juvenile rats. Neuroscience. 230, 114-120.

Moller, A.R., 1983, Auditory Physiology. Academic Press, Toronto.

Page 135: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

124

Moore, D.R., Aitkin, L.M., 1975. Rearing in an acoustically unusual environment -

effects on neural auditory responses. Neurosci Lett. 1, 29-34.

Morest, D.K., Oliver, D.L., 1984. The neuronal architecture of the inferior colliculus in

the cat: Defining the functional anatomy of the auditory midbrain. J Comp Neurol. 222,

209-236.

Morgan, J.I., Curran, T., 1991. Stimulus-transcription coupling in the nervous system:

Involvement of the inducible proto-oncogenes fos and jun. Annu Rev Neurosci. 14, 421-

451.

Nakahara, H., Zhang, L.I., Merzenich, M.M., 2004. Specialization of primary auditory

cortex processing by sound exposure in the "critical period". Proc Natl Acad Sci U S A.

101, 7170-7174.

Nakamura, M., Rosahl, S.K., Alkahlout, E., Walter, G.F., Samii, M.M., 2005. Electrical

stimulation of the cochlear nerve in rats: Analysis of c-fos expression in auditory

brainstem nuclei. Brain Res. 1031, 39-55.

Norena, A.J., Gourevitch, B., Aizawa, N., Eggermont, J.J., 2006. Spectrally enhanced

acoustic environment disrupts frequency representation in cat auditory cortex. Nat

Neurosci. 9, 932-939.

Oertel, D., Wu, S., Hirsch, J.A., 1988. Electrical characteristics of cells and neuronal

circuitry in the cochlear nuclei studied with intracellular recordings from brain slices. In

Auditory Function: Neurobiological Bases of Hearing. G.M. Edelman, W.E. Gall, W.M.

Cowan, eds. Wiley, New York, pp. 313-336.

Ojima, H., Honda, C.N., Jones, E.G., 1992. Characteristics of intracellularly injected

infragranular pyramidal neurons in cat primary auditory cortex. Cereb Cortex. 2, 197-

216.

Oliver, D.L., Izquierdo, M.A., Malmierca, M.S., 2011. Persistent effects of early

augmented acoustic environment on the auditory brainstem. Neuroscience. 184, 75-87.

Oliver, D.L., Winer, J.A., Beckius, G.E., Saint Marie, R.L., 1994. Morphology of

GABAergic neurons in the inferior colliculus of the cat. J Comp Neurol. 340, 27-42.

Oliver, D.L., Morest, D.K., 1984. The central nucleus of the inferior colliculus in the cat.

J Comp Neurol. 222, 237-264.

Palombi, P.S., Caspary, D.M., 1996a. GABA inputs control discharge rate primarily

within frequency receptive fields of inferior colliculus neurons. J Neurophysiol. 75, 2211-

2219.

Page 136: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

125

Palombi, P.S., Caspary, D.M., 1996b. Physiology of the young adult fischer 344 rat

inferior colliculus: Responses to contralateral monaural stimuli. Hear Res. 100, 41-58.

Paxinos, G., Watson, C., 1982, The Rat Brain in Stereotaxic Coordinates. Academic

Press, New York, USA.

Petersohn, D., Thiel, G., 1996. Role of zinc-finger proteins Sp1 and zif268/egr-1 in

transcriptional regulation of the human synaptobrevin II gene. Eur J Biochem. 239, 827-

834.

Petersohn, D., Schoch, S., Brinkmann, D.R., Thiel, G., 1995. The human synapsin II gene

promoter. possible role for the transcription factor zif268/egr-1, polyoma enhancer

activator 3, and AP2. J Biol Chem. 270, 24361-24369.

Pienkowski, M., Eggermont, J.J., 2011. Cortical tonotopic map plasticity and behavior.

Neurosci Biobehav Rev. 35, 2117-2128.

Pienkowski, M., Eggermont, J.J., 2010. Passive exposure of adult cats to moderate-level

tone pip ensembles differentially decreases AI and AII responsiveness in the exposure

frequency range. Hear Res. 268, 151-162.

Pienkowski, M., Eggermont, J.J., 2009. Long-term, partially-reversible reorganization of

frequency tuning in mature cat primary auditory cortex can be induced by passive

exposure to moderate-level sounds. Hear Res. 257, 24-40.

Pienkowski, M., Harrison, R.V., 2005a. Tone frequency maps and receptive fields in the

developing chinchilla auditory cortex. J Neurophysiol. 93, 454-466.

Pienkowski, M., Harrison, R.V., 2005b. Tone responses in core versus belt auditory

cortex in the developing chinchilla. J Comp Neurol. 492, 101-109.

Platenik, J., Kuramoto, N., Yoneda, Y., 2000. Molecular mechanisms associated with

long-term consolidation of the NMDA signals. Life Sci. 67, 335-364.

Pollak, G.D., Burger, R.M., Klug, A., 2003. Dissecting the circuitry of the auditory

system. Trends Neurosci. 26, 33-39.

Pollak, G.D., Park, T.J., 1993. The effects of GABAergic inhibition on monaural

response properties of neurons in the mustache bat's inferior colliculus. Hear Res. 65, 99-

117.

Poon, P.W., Chen, X., 1992. Postnatal exposure to tones alters the tuning characteristics

of inferior collicular neurons in the rat. Brain Res. 585, 391-394.

Page 137: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

126

Pospelov, V.A., Pospelova, T.V., Julien, J.P., 1994. AP-1 and krox-24 transcription

factors activate the neurofilament light gene promoter in P19 embryonal carcinoma cells.

Cell Growth Differ. 5, 187-196.

Preston, G.A., Lyon, T.T., Yin, Y., Lang, J.E., Solomon, G., Annab, L., Srinivasan, D.G.,

Alcorta, D.A., Barrett, J.C., 1996. Induction of apoptosis by c-fos protein. Mol Cell Biol.

16, 211-218.

Prieto, J.J., Winer, J.A., 1999. Layer VI in cat primary auditory cortex: Golgi study and

sublaminar origins of projection neurons. J Comp Neurol. 404, 332-358.

Purves, D., Augustine, G.J., Fitzpatrick, D., Katz, L.C., LaMantia, A., McNamara, J.O.,

1997, Neuroscience. Sinauer Associates, Inc., Sunderland, Massachusetts.

Qian, Y., Jen, P.H., 1994. Fos-like immunoreactivity elicited by sound stimulation in the

auditory neurons of the big brown bat eptesicus fuscus. Brain Res. 664, 241-246.

Querleu, D., Renard, X., Versyp, F., Paris-Delrue, L., Crepin, G., 1988. Fetal hearing.

Eur J Obstet Gynecol Reprod Biol. 28, 191-212.

Quinlan, E.M., Philpot, B.D., Huganir, R.L., Bear, M.F., 1999. Rapid, experience-

dependent expression of synaptic NMDA receptors in visual cortex in vivo. Nat

Neurosci. 2, 352-357.

Rajan, R., Irvine, D.R., 1998. Absence of plasticity of the frequency map in dorsal

cochlear nucleus of adult cats after unilateral partial cochlear lesions. J Comp Neurol.

399, 35-46.

Rasmusson, D.D., 1982. Reorganization of raccoon somatosensory cortex following

removal of the fifth digit. J Comp Neurol. 205, 313-326.

Recanzone, G.H., Schreiner, C.E., Merzenich, M.M., 1993. Plasticity in the frequency

representation of primary auditory cortex following discrimination training in adult owl

monkeys. J Neurosci. 13, 87-103.

Reimer, K., 1993. Simultaneous demonstration of fos-like immunoreactivity and 2-

deoxy-glucose uptake in the inferior colliculus of the mouse. Brain Res. 616, 339-343.

Robertson, D., Irvine, D.R., 1989. Plasticity of frequency organization in auditory cortex

of guinea pigs with partial unilateral deafness. J Comp Neurol. 282, 456-471.

Rosen, K.M., McCormack, M.A., Villa-Komaroff, L., Mower, G.D., 1992. Brief visual

experience induces immediate early gene expression in the cat visual cortex. Proc Natl

Acad Sci U S A. 89, 5437-5441.

Page 138: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

127

Roux, P., Blanchard, J.M., Fernandez, A., Lamb, N., Jeanteur, P., Piechaczyk, M., 1990.

Nuclear localization of c-fos, but not v-fos proteins, is controlled by extracellular signals.

Cell. 63, 341-351.

Saint Marie, R.L., Shneiderman, A., Stanforth, D.A., 1997. Patterns of gamma-

aminobutyric acid and glycine immunoreactivities reflect structural and functional

differences of the cat lateral lemniscal nuclei. J Comp Neurol. 389, 264-276.

Sallaz, M., Jourdan, F., 1993. C-fos expression and 2-deoxyglucose uptake in the

olfactory bulb of odour-stimulated awake rats. Neuroreport. 4, 55-58.

Sanes, D.H., Constantine-Paton, M., 1985. The sharpening of frequency tuning curves

requires patterned activity during development in the mouse, mus musculus. J Neurosci.

5, 1152-1166.

Scheich, H., Zuschratter, W., 1995. Mapping of stimulus features and meaning in gerbil

auditory cortex with 2-deoxyglucose and c-fos antibodies. Behav Brain Res. 66, 195-205.

Schreiner, C.E., Polley, D.B., 2014. Auditory map plasticity: Diversity in causes and

consequences. Curr Opin Neurobiol. 24, 143-156.

Schreiner, C.E., Langner, G., 1997. Laminar fine structure of frequency organization in

auditory midbrain. Nature. 388, 383-386.

Sheng, M., Cummings, J., Roldan, L.A., Jan, Y.N., Jan, L.Y., 1994. Changing subunit

composition of heteromeric NMDA receptors during development of rat cortex. Nature.

368, 144-147.

Sheng, M., Greenberg, M.E., 1990. The regulation and function of c-fos and other

immediate early genes in the nervous system. Neuron. 4, 477-485.

Shinonaga, Y., Takada, M., Mizuno, N., 1994. Direct projections from the non-laminated

divisions of the medial geniculate nucleus to the temporal polar cortex and amygdala in

the cat. J Comp Neurol. 340, 405-426.

Silverthorn, D.U., 2004, Human Physiology: An Integrated Approach. Pearson Education

Inc., San Francisco, California.

Stanton, S.G., Harrison, R.V., 1996. Abnormal cochleotopic organization in the auditory

cortex of cats reared in a frequency augmented environment. Audit Neurosci. 2, 97-107.

Stanton, S.G., Harrison, R.V., 2000. Projections from the medial geniculate body to

primary auditory cortex in neonatally deafened cats. J Comp Neurol. 426, 117-129.

Steward, O., Worley, P.F., 2001a. A cellular mechanism for targeting newly synthesized

mRNAs to synaptic sites on dendrites. Proc Natl Acad Sci U S A. 98, 7062-7068.

Page 139: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

128

Steward, O., Worley, P.F., 2001b. Selective targeting of newly synthesized arc mRNA to

active synapses requires NMDA receptor activation. Neuron. 30, 227-240.

Steward, O., Wallace, C.S., Lyford, G.L., Worley, P.F., 1998. Synaptic activation causes

the mRNA for the IEG arc to localize selectively near activated postsynaptic sites on

dendrites. Neuron. 21, 741-751.

Syka, J., Popelar, J., Kvasnak, E., Astl, J., 2000. Response properties of neurons in the

central nucleus and external and dorsal cortices of the inferior colliculus in guinea pig.

Exp Brain Res. 133, 254-266.

Thiel, G., Schoch, S., Petersohn, D., 1994. Regulation of synapsin I gene expression by

the zinc finger transcription factor zif268/egr-1. J Biol Chem. 269, 15294-15301.

Turrigiano, G.G., Leslie, K.R., Desai, N.S., Rutherford, L.C., Nelson, S.B., 1998.

Activity-dependent scaling of quantal amplitude in neocortical neurons. Nature. 391,

892-896.

von Békésy, G., 1960, Experiments in Hearing. McGraw-Hill, New York.

Wallhausser, E., Scheich, H., 1987. Auditory imprinting leads to differential 2-

deoxyglucose uptake and dendritic spine loss in the chick rostral forebrain. Brain Res.

428, 29-44.

Wiesel, T.N., 1982. Postnatal development of the visual cortex and the influence of

environment. Nature. 299, 583-591.

Wiesel, T.N., Hubel, D.H., 1965. Comparison of the effects of unilateral and bilateral eye

closure on cortical unit responses in kittens. J Neurophysiol. 28, 1029-1040.

Wiesel, T.N., Hubel, D.H., 1963. Single-cell responses in striate cortex of kittens

deprived of vision in one eye. J Neurophysiol. 26, 1003-1017.

Winer, J.A., 2005. Three systems of descending projections to the inferior colliculus. In

The Inferior Colliculus. J.A. Winer, C.E. Schreiner, eds. Springer, Jeffery A. Winer,

Christoph E. Schreiner, pp. 231-247.

Winer, J.A., Schreiner, C.E., 2005. The central auditory system. In The Inferior

Colliculus. J.A. Winer, C.E. Schreiner, eds. Springer, Berlin Heidelberg New York, pp.

1-68.

Winer, J.A., Larue, D.T., Diehl, J.J., Hefti, B.J., 1998. Auditory cortical projections to the

cat inferior colliculus. J Comp Neurol. 400, 147-174.

Page 140: Developmental Plasticity of Tonotopic Maps in …...ii Abstract Developmental Plasticity of Tonotopic Maps in Chinchilla Auditory Midbrain Lisa Margaret D’Alessandro Doctor of Philosophy

129

Winer, J.A., Saint Marie, R.L., Larue, D.T., Oliver, D.L., 1996. GABAergic feedforward

projections from the inferior colliculus to the medial geniculate body. Proc Natl Acad Sci

U S A. 93, 8005-8010.

Wu, G., Malinow, R., Cline, H.T., 1996. Maturation of a central glutamatergic synapse.

Science. 274, 972-976.

Xie, R., Meitzen, J., Pollak, G.D., 2005. Differing roles of inhibition in hierarchical

processing of species-specific calls in auditory brainstem nuclei. J Neurophysiol. 94,

4019-4037.

Yu, X., Sanes, D.H., Aristizabal, O., Wadghiri, Y.Z., Turnbull, D.H., 2007. Large-scale

reorganization of the tonotopic map in mouse auditory midbrain revealed by MRI. Proc

Natl Acad Sci U S A. 104, 12193-12198.

Yuste, R., Bonhoeffer, T., 2001. Morphological changes in dendritic spines associated

with long-term synaptic plasticity. Annu Rev Neurosci. 24, 1071-1089.

Zhang, L.I., Bao, S., Merzenich, M.M., 2001. Persistent and specific influences of early

acoustic environments on primary auditory cortex. Nat Neurosci. 4, 1123-1130.

Zhang, Y., Wu, S.H., 2000. Long-term potentiation in the inferior colliculus studied in rat

brain slice. Hear Res. 147, 92-103.

Zhou, X., Panizzutti, R., de Villers-Sidani, E., Madeira, C., Merzenich, M.M., 2011.

Natural restoration of critical period plasticity in the juvenile and adult primary auditory

cortex. J Neurosci. 31, 5625-5634.