You are on page 1of 32

Provided for non-commercial research and educational use.

Not for reproduction, distribution or commercial use.

This article was originally published in the The Senses: A Comprehensive Reference,
published by Elsevier, and the attached copy is provided by Elsevier for the author’s benefit
and for the benefit of the author’s institution, for non-commercial research and educational
use including without limitation use in instruction at your institution,

sending it to specific colleagues who you know, and providing a copy to your institution’s
administrator.

All other uses, reproduction and distribution, including without limitation commercial reprints,
selling or licensing copies or access, or posting on open internet sites, your personal or
institution’s website or repository, are prohibited. For exceptions, permission may be sought
for such use through Elsevier’s permissions site at:

http://www.elsevier.com/locate/permissionusematerial

M A Rutherford and W M Roberts, Afferent Synaptic Mechanisms. In: Allan I. Basbaum,


Akimichi Kaneko, Gordon M. Shepherd and Gerald Westheimer, editors The Senses:
A Comprehensive Reference, Vol 3, Audition, Peter Dallos and
Donata Oertel. San Diego: Academic Press; 2008. p. 365-396.
Author's personal copy

3.22 Afferent Synaptic Mechanisms


M A Rutherford and W M Roberts, University of Oregon, Eugene, OR, USA
ª 2008 Elsevier Inc. All rights reserved.

3.22.1 Introduction 366


3.22.2 Afferent Neurons of Acousticolateralis Sensory Systems 366
3.22.2.1 Spiral Ganglion Neurons 367
3.22.2.2 Auditory Nerve Fibers 368
3.22.2.3 Innervation Patterns and the Function of Auditory Organs 371
3.22.3 Hair Cells 372
3.22.3.1 Hair Cell Shape in Relation to Synaptic Signaling 372
3.22.3.2 The Synaptic Body and Synaptic Vesicle Pools 372
3.22.3.3 Calcium Buffering in Hair Cells 375
3.22.4 Ribbon Synapses – A Small but Diverse Class 377
3.22.4.1 The Synaptic Body (Ribbon) and Its Vesicles 377
3.22.4.2 Molecular Specializations of Ribbon Synapses 378
3.22.4.3 Anatomy of the Synaptic Body: Heterogeneity of Structure 380
3.22.5 The Synaptic Vesicle Cycle in Hair Cells 381
3.22.5.1 Endocytosis 381
3.22.5.2 Vesicle Mobility 381
3.22.5.3 Docking and Priming 382
3.22.5.4 Transmitter Release and Clearance 382
3.22.5.5 Relating Anatomical and Functional Vesicle Pools 383
3.22.6 Physiology of the Hair Cell Afferent Synapse 384
3.22.6.1 Encoding of Phase 385
3.22.6.2 Encoding of Frequency 386
3.22.6.3 Baseline and Evoked Synaptic Activity 386
3.22.6.4 Tuning Mechanisms in Hair Cells 388
3.22.6.4.1 Frequency-selective augmentation of membrane potential oscillation 388
3.22.6.4.2 Synaptic frequency selectivity 388
3.22.7 Auditory Nerve Fiber Firing Patterns Mirror Presynaptic Vesicle Pool Dynamics 389
3.22.7.1 Adaptation of the Auditory Nerve Response 389
3.22.7.2 The Role of the Synaptic Body in Auditory Nerve Fiber Spike Timing 390
3.22.8 Synapse to Psychophysics 391
References 391

Glossary
active zone (AZ) The area where synaptic vesicles characteristic frequency (CF) The frequency of
fuse with the presynaptic membrane. The AZ is maximum sensitivity to sound.
juxtaposed to the postsynaptic terminal across the cytomatrix at the active zone (CAZ) The cytos-
synaptic cleft. keletal meshwork comprised of multidomain
afferent The flow of signals from the periphery proteins that aggregate through specific protein–
toward the central nervous system (CNS). protein and protein–lipid interactions. The CAZ
auditory nerve fiber (ANF) The axon of an audi- creates a scaffold upon which the presynaptic
tory afferent neuron that projects to the brainstem machinery is assembled.
in the vestibulocochlear (eighth cranial) nerve. In hair cells The sensory receptors of the auditory
mammals, ANFs have cell bodies in the spiral system and related sense organs of the ear and
ganglion. lateral line of vertebrate animals.

365
The Senses: A Comprehensive Reference, vol. 3, pp. 365-395
Author's personal copy
366 Afferent Synaptic Mechanisms

inner hair cell (IHC) One of the highly innervated agreement that the SB serves an important pur-
hair cells in the mammalian ear that are primarily pose, which is not presently known. It harbors
responsible for the perception of sound. They are synaptic vesicles attached to it via fine filaments
located in a single row near the modiolar (inner) that are thought to either facilitate or impede the
edge of the organ of Corti. movement of vesicles along its surface.
root mean square (RMS) A type of average com- synaptic vesicle (SV) Spherical lipid bilayer that
puted by first squaring a set of values, then contains neurotransmitter.
computing the mean of the squared values, then synaptic vesicle cycle The recurrent processes
taking the square root of the mean. The RMS by which SVs are continuously created, used and
average deviation of N samples from their mean is recreated. Exocytosis of the neurotransmitter is
p
(N  1) times the standard error of the mean. triggered by calcium and involves fusion of the SV
sound pressure level (SPL) The RMS pressure membrane with the plasmalemma and/or the for-
change during a sound, expressed on the logarith- mation of a transient aqueous connection between
mic decibel (dB) scale, measured in the vicinity of the SV and the extracellular space. After exocyto-
the listener during a specified time period. sis, SVs are reformed, refilled, and reenter the
spiral ganglion The cell bodies of the afferent exocytic pathway.
neurons that innervate the organ of Corti; housed in threshold The SPL needed to produce a criterion
the bony spiral lamina of the mammalian cochlea. response. Defined in this way, threshold is a mea-
spontaneous rate (SR) Frequency of spikes in an sure of sensitivity to sound: low threshold means
auditory nerve fiber in the absence of an applied high sensitivity. Sensitivity is usually expressed in
sound stimulus. units of response/stimulus, but in auditory physiol-
synaptic body or ribbon (SB) The most obvious ogy threshold is traditionally defined in units of
presynaptic structure at hair cell afferent synapses, stimulus only, omitting explicit reference to the
and the subject of many carefully designed experi- response criterion used in a particular experiment.
ments to reveal its function(s). There is general

3.22.1 Introduction 3.22.2 Afferent Neurons of


Acousticolateralis Sensory Systems
We will begin with an overview of the hair cell
afferent synapse and related structures, with a focus The primary afferent neurons in the auditory and
on mammalian inner hair cells (IHCs) and spiral vestibular organs of all vertebrate species, and in
ganglion neurons. Our review will then turn to elec- related mechanosensory and electrosensory organs
trophysiological studies and discussions of the of the lateral line system of aquatic vertebrates are
specializations that may endow these first auditory bipolar neurons that have cell bodies located in or
synapses with their ability to maintain fast and reli- near the ear. They extend one process peripherally to
able, information-rich signaling to the brain. Readers the sensory epithelium, the other centrally to the
unfamiliar with the basics of cellular neuroscience brainstem. The peripheral processes have often
will better appreciate the unique properties of audi- been called dendrites, but most are myelinated and
tory afferent synapses after reviewing the properties of can properly be called axons. This distinction is
conventional chemical synapses described in the many likely to be important for the assembly and mainte-
excellent neuroscience textbooks (e.g., Delcomyn, F., nance of afferent synapses because axons and
1998; Kandel, E. R. et al., 2000; Hille, B., 2001; dendrites use different protein-targeting signals
Nicholls, J. G. et al., 2001). Additional information spe- (Burack, M. A. et al., 2000). The afferent neurons in
cific to hair cell afferent synapses can be found in recent auditory and vestibular organs have cell bodies
literature reviews (Nouvian, R. et al., 2006; Moser, T. located close to the corresponding sensory epithe-
et al., 2006a; 2006b). lium, and therefore have short peripheral axons.

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
Afferent Synaptic Mechanisms 367

Lateral line afferent neurons have long peripheral cochlea thus represents a case of synaptic divergence
axons that extend from cell bodies in the lateral line in which each IHC controls many private communi-
ganglia near the ear to sensory neuromasts on the cation channels to the brain. The function of this
head and trunk (Dambly-Chaudiere, C. et al., 2003). unusual arrangement is only partially understood,
but it is easy to speculate that the dynamic membrane
potential of the hair cell contains more information
3.22.2.1 Spiral Ganglion Neurons
than can be transmitted across one synapse and
In mammals, the cell bodies of auditory afferent encoded in a single spike train.
neurons reside in the spiral ganglion, which parallels Information is degraded by events that alter a
the inner (modiolar) edge of the organ of Corti signal in an unpredictable way (noise). The synaptic
(Figure 1(a)). Type I spiral ganglion neurons are noise introduced by the stochastic nature of exocy-
responsive to acoustic stimulation and constitute tosis can be reduced by having synaptic currents from
90–95% of cochlear nerve afferents (Spoendlin, H., many AZs converge on the spike initiation zone of
1969; Liberman, M. C., 1982). Intracellular labeling one afferent axon, as occurs at the calyx synapses of
with horseradish peroxidase and subsequent histolo- vestibular hair cells (Goldberg, J. M., 1996); but the
gical processing revealed that they contact only the mammalian auditory system uses a different strategy
IHCs (Robertson, D., 1984). Both the peripheral and in which the maximum possible separation of infor-
the central axons of a type I afferent neuron are mation channels to the brain is maintained by having
myelinated and have a cluster of voltage-gated each synaptic AZ control action potential generation
Naþ channels (Nav 1.6) at each node of Ranvier. in a separate afferent axon. If separate information
Naþ channels are also found in the unmyelinated channels later converge onto a single higher-order
axon segment between the postsynaptic terminal neuron in the brain stem (e.g., a globular bushy cell in
and the most distal peripheral node, suggesting that the cochlear nucleus that detects coincident input),
spikes may be initiated at or near the site of synaptic this arrangement can reduce the effects of variability
input (Hossain, W. A. et al., 2005), as occurs in verte- in the spike conduction time from the ear to the
brate skeletal muscle (Kandel, E. R. et al., 2000). central nervous system (CNS) and other noise
Minimizing the distance from the synapse to the sources that are downstream of the afferent synapse.
spike initiation site is expected to minimize attenua- A small population of afferent axons in the mam-
tion and low-pass filtering of excitatory postsynaptic malian cochlea (type II; Figure 1(a)) are
potentials (EPSPs). The cell bodies of type I spiral unmyelinated (Liberman, M. C. et al., 1990). Like
ganglion neurons are also wrapped in a loose myelin- the type I afferents, they are labeled by antibodies
like sheath that may speed conduction, reduce tem- against Nav 1.6 and may initiate action potentials
poral jitter, or decrease the probability of conduction near their distal tips (Hossain, W. A. et al., 2005).
failure across the cell body during high-frequency These outer hair cell (OHC) afferents are thought
firing (Hossain, W. A. et al., 2005). to monitor the state of the organ of Corti, but not to
In mammals, the peripheral axons of type I affer- contribute directly to the perception of sound, the
ents (also known as radial fibers) remain unbranched role ascribed to the IHCs. Very little is known about
and contact only a single IHC. Unlike most neurons, the physiology of type II afferents, and they will not
which receive synaptic inputs from hundreds or be considered further in this chapter.
thousands of separate sites of synaptic contact (active Although the peripheral neural network in the ear
zones; AZs), each type I afferent neuron typically is less elaborate than that of the retina and contains
receives sensory input from a single hair cell AZ far fewer cell types, electron micrographs of the
(Spoendlin, H., 1969; Liberman, M. C., 1980; organ of Corti do show several additional types of
Spoendlin, H., 1985; Hashimoto, S. et al., 1990). In synapses that could provide important functional
other vertebrates, most auditory afferent neurons interactions, including local feedback (Raphael, Y.
make multiple postsynaptic contacts on one or more and Altschuler, R. A., 2003; Sobkowicz, H. M. et al.,
hair cells (Lewis, E. R. et al., 1982). 2003). These include connections between IHCs and
Each mammalian IHC provides the sole input to OHCs via afferent and efferent fibers that contact
5–30 type I afferent axons, which vary in number by both hair cell types (Sobkowicz, H. M. et al., 2004).
species and hair cell position on the cochlear tonoto- Nothing is known about the function of these
pic axis (Kiang, N. Y. S. et al., 1982; Liberman, M. C., synapses in auditory signal processing. Efferent
1982; Fuchs, P. A. et al., 2003). The mammalian synapses are typically seen on the postsynaptic

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
368 Afferent Synaptic Mechanisms

(a)
Spiral
lamina
Type l Node of Tectorial membrane
Ranvier
IHC
Type ll
OHCs

Bas
ilar
(b) (c) mem
Endolymph bran
e
Perilymph Tight
SB junction
10 mV
Modiolar Pillar 2 ms

EPSPs

0.1 MET
mV Tone
Afferent
bouton

Figure 1 Auditory afferent synapses. (a) Schematic cross section through the mammalian organ of Corti, showing the
bipolar cell bodies of afferent neurons (left), which each send one axon peripherally toward the hair cells and another axon
centrally to the brainstem. In mammals, most afferent axons (type I) are myelinated (green), unbranched, and contact a single
inner hair cell (IHC). A smaller population (type II) are unmyelinated and branch to contact several outer hair cells. A high density
of voltage-gated sodium channels (Nav 1.6, red) are located at nodes of Ranvier, on the unmyelinated distal axon segments of
all afferent neurons, and at both axon hillocks of type II neurons. (b) Each IHC supplies the only input to tens of type I afferent
neurons at active zones. Large-diameter afferent axons usually contact the pillar face of the IHC, whereas smaller-diameter
axons contact the modiolar side. (c) Electrophysiological recording of afferent synaptic activity in vivo. A 200 Hz sinusoidal
sound delivered to the goldfish sacculus (bottom trace) evoked an extracellular field potential (middle trace), analogous to a
cochlear microphonic potential (see Chapter Hair Cell Transduction and Adaptation: Physiology and Molecular Mechanisms),
that is primarily a measure of mechanoelectrical transduction (MET). The top trace is an intracellular recording from an afferent
axon, in which spikes had been blocked to reveal the EPSPs, which are synchronized to the sound waveform and show strong
synaptic depression. (c) Adapted from Furukawa, T. and Matsuura, S. 1978. Adaptive rundown of excitatory postsynaptic
potentials at synapses between hair cells and eighth nerve fibres in the goldfish. J. Physiol. 276, 193–209, with permission from
Blackwell Publishing.

terminals of mammalian type I afferents, where they more numerous than the axons of type II neurons
may modulate synaptic transmission from the hair (Liberman, M. C., 1982).
cell to the afferent axon. Efferent axons also make In vivo recordings from single ANFs show that
inhibitory cholinergic synapses directly onto hair nearly all fire action potentials spontaneously (i.e., in
cells in most vertebrate auditory and vestibular a quiet laboratory environment with no experimen-
organs, including mammalian IHCs and OHCs (see tally imposed stimulus). The baseline or spontaneous
Chapter Efferent System). firing rate (SR) in some ANFs exceeds 100 spikes s1,
while others fire fewer than 5 spikes s1, with no
apparent correlation between a fiber’s characteristic
frequency (CF) and SR (Kiang, N. Y. S. et al., 1965;
3.22.2.2 Auditory Nerve Fibers
Tsuji, J. and Liberman, M. C., 1997). CF varies with
The central axons of spiral ganglion neurons project position along the tonotopic axis whereas SR has been
to the brainstem in the eighth cranial nerve, where correlated with both axon diameter and subcellular
they are called auditory nerve fibers (ANFs). Nearly synaptic structure in the guinea-pig and cat
all recordings of action potentials in mammalian (Liberman, M. C., 1980; Merchan-Perez, A. and
ANFs are presumed to be from axons of type I spiral Liberman, M. C., 1996; Tsuji, J. and Liberman, M. C.,
ganglion neurons because they are larger and much 1997). In these studies of afferent structure/function,

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
Afferent Synaptic Mechanisms 369

intracellular recording and labeling of individual A threshold nonlinearity in the peripheral audi-
ANFs followed by serial-section electron microscopy tory system would be detrimental to subsequent
(EM) showed that ANFs of high SR belong to spiral neural processing because it would eliminate infor-
ganglion neurons that have large diameter peripheral mation that could otherwise be recovered by
axons and contact IHCs predominantly on the pillar temporal averaging (integration) or filtering, which
face (Figure 1(b)); however, this segregation of axonal can in principle pull an arbitrarily small periodic
diameters around the hair cell circumference was not signal out of aperiodic background noise. This ability
observed in one study of the gerbil (Slepecky, N. B. is crucial for systems designed to detect small noisy
et al., 2000). Low- and intermediate-SR fibers of the cat signals. For example, a sinusoidal signal that is super-
and guinea-pig contact the modiolar face and have imposed on additive white noise can be passed
synapses with larger presynaptic dense bodies (called through a narrow-band linear filter that leaves the
synaptic bodies, SBs, or ribbons) and more vesicles signal unchanged while reducing the noise in propor-
than synapses with high-SR fibers. Because the 5–30 tion to the bandwidth of the filter. Narrowing the
afferent synapses on each IHC display this range of filter’s bandwidth can thus increase the signal-to-
ultrastructural characteristics and modiolar/pillar noise ratio as much needed to detect an arbitrarily
position, and because ANFs of the same CF differ in small periodic signal, albeit at the expense of tem-
SR, it is inferred that each IHC synapses onto both poral resolution (Bracewell, R. N., 1999). The
high-SR fibers and intermediate-/low-SR fibers. auditory system’s use of temporal integration to
To quantify the strength of a sound, most experi- improve signal-to-noise ratio is seen experimentally as
menters report the sound pressure level (SPL), an inverse relation between threshold and sound dura-
expressed on the logarithmic decibel scale (dB). tion ( Jeffress, L. A., 1968; Heil, P. and Neubauer, H.,
Each 20 dB increase in SPL corresponds to a tenfold 2001; 2003). The observation that sounds too soft to
be audible can nevertheless influence the audibility
multiplication of root mean square (RMS) sound
of subsequent sounds (Plack, C. J. et al., 2006) supports
pressure. The zero point (0 dB) is often defined as
the idea that subthreshold sounds still produce some
20 mPa, which is approximately the amplitude of a
response.
barely perceptible sound at frequencies where human
The peripheral auditory system eliminates any
hearing is most sensitive (2–5 kHz). Compared to the
threshold cut-off in the afferent response to faint
standard atmospheric pressure of 100 kPa, 20 mPa
sounds by eliminating all slack from the system,
amounts to a change of only about 1 part in 5 billion.
ensuring that arbitrarily weak sounds evoke a change
An ANF’s sensitivity to sound (Figure 2) is typi-
in ANF firing rate that is proportional to the sound
cally measured as the SPL needed to increase its
amplitude, at least for low-frequency tones. This
firing rate by a criterion number of spikes per second
begins with mechanosensory transduction, which
above the SR, or alternatively, to produce a criterion
uses feedback to maintain partial activation of the
increase in ANF discharge synchrony with the sound transduction channels even in complete silence.
stimulus (Rose, J. E. et al., 1967; 1971; Johnson, D. H., The resulting resting transduction current places
1980). This measure, which is often called the fiber’s the hair cell’s resting membrane potential in a range
threshold, depends on the response criterion, stimu- where the presynaptic calcium channels are also par-
lus duration, and other factors that are specific to the tially activated (Koschak, A. et al., 2001; Xu, W. and
experimental protocol (Heil, P. and Neubauer, H., Lipscombe, D., 2001; Johnson, S. L. et al., 2005). This
2003). The term threshold is typically used in biology steady-state activation of the calcium current (i.e., the
to mean a stimulus level below which there is no random, thermally driven, independent gating of
response (e.g., the threshold for generating an action individual calcium channels that occurs at constant
potential), and therefore implies the existence of a membrane potential) may be sufficient to catalyze
strong nonlinearity in the stimulus–response relation baseline transmitter release (Brandt, A. et al., 2005)
(a cut-off below which there is no response). The use that drives the so-called spontaneous spiking in
of the term threshold in auditory physiology is quite ANFs (Liberman, M. C. and Kiang, N. Y. S., 1978;
different. It carries no implication of a threshold Sewell, W. F., 1984; Glowatzki, E. and Fuchs, P. A.,
nonlinearity, only that the response has fallen below 2002; Keen, E. C. and Hudspeth, A. J., 2006). Thus,
some criterion level. In fact, the initial stages of each stage of processing is maintained in a partially
auditory processing are constructed to avoid thresh- active state even in complete silence, thereby avoid-
old nonlinearities. ing a threshold nonlinearity and the resulting loss of

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
370 Afferent Synaptic Mechanisms

(a) 120 (b)

110 120

100 100
Threshold (dB SPL)

Threshold (dB SPL)


90 80

80 60

70 40

60 20

50 0
0.1 1
Frequency (kHz)
40
0.01 0.1 1
100 (c)
1
90
0.8
80 0.6
Threshold (dB SPL)

Vector strength
70
0.4
60

50
0.2
40

30
0.1
0.01 0.1 1
20 Frequency (kHz)
0.1 1 10
Frequency (kHz)
Figure 2 Tuning curves and phase locking in avian (emu) auditory nerve fibers (ANFs). (a) The sound pressure level (SPL)
required to evoke a criterion response, in this case, a 20 Hz increase in ANF firing rate above the spontaneous rate (SR),
plotted as a function of the frequency of the tonal stimulus. This type of measurement is often called the fiber’s threshold,
although it does not mean that the fiber fails to respond to less intense sounds, only that the response falls below the
experimenter’s criterion. Data from two ANFs are shown. Each panel shows one fiber’s tuning curve as a continuous line and
phase histograms (insets) from the same fiber, each spanning a time frame of two cycles on the x-axis and displaying the
probability of a spike in particular time bins within the cycle. The stimulus for the insets was at the frequency-level
combinations indicated by the associated filled, inverted triangles. The phase relationship of spike occurrence to the period of
the tonal stimulus changes as a function of frequency, but at the characteristic frequency (the low point on the curve) the
phase relationship does not vary with changes in sound intensity (dB SPL). (b) Tuning curves for four ANFs of different
characteristic frequency. (c) When vector strength (a measure of the degree of phase locking, on a scale of 0 to 1, derived
from phase histograms like those shown in (a)) is plotted as a function of stimulus frequency around characteristic frequency
for 73 different afferent fibers, it is apparent that cochlear ANFs in the emu exhibit a sharp decline in phase-locked firing at
frequencies above 1 kHz. Reprinted with permission from Manley, G. A., Köppl, C., and Yates, G. K., J. Acoust. Soc. Am.,
Vol. 101, pages 1560-1573, 1997. Copyright 1997, American Institute of Physics.

information about small signals. Because several steps high sensitivity to small sounds (Indresano, A. A. et al.,
in auditory reception involve discrete all-or-none 2004). This phenomenon has been studied in many
events (e.g., the opening of individual ion channels, physical systems under the odd name of stochastic
the fusion of individual synaptic vesicles (SVs) with resonance.
the plasmalemma, and the generation of individual At frequencies above a few kilohertz, where sound
action potentials in ANFs), the presence of thermally perception relies on nonlinear transduction at the
driven noise in the system is crucial to maintaining foot of the stimulus–response curve to produce the

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
Afferent Synaptic Mechanisms 371

DC component of the receptor potential (see Section implying a mean closed time of <10 ms. If calcium
3.22.6), a nonlinear response to small sounds may be channels in IHCs have properties similar to these,
unavoidable. then each channel is expected to generate >100
Fibers that are the most sensitive to sound also openings per second at the resting potential, and the
have the highest SR. The equivalent statement in ensemble of 100 calcium channels at a typical AZ
terms of threshold is that SR varies inversely with are expected to generate >104 openings per second.
threshold. In cats, ANFs have been divided into three Given that the highest SRs in ANFs are 100 s1,
classes based on SR: low-SR fibers (<0.5 Hz) have the fewer than 1% of calcium channel openings are pre-
highest threshold (20–60 dB higher than high-SR dicted to trigger exocytosis.
fibers), high-SR fibers (18 Hz) have the lowest
thresholds, and medium-SR fibers have an inter-
3.22.2.3 Innervation Patterns and the
mediate sensitivity (Merchan-Perez, A. and
Function of Auditory Organs
Liberman, M. C., 1996). Because ANFs with the
same CF can have high-, intermediate-, or low-SR, The pattern of afferent innervation onto hair cells
it seems likely that a single IHC forms synapses with contributes greatly to the response properties
representatives of all three functional groups. observed in afferent fiber recordings. To illustrate
Unlike many types of neurons that fire either with this, we will consider two very different auditory
a fairly constant interspike interval or in bursts, the organs of the inner ear: the mammalian cochlea and
exponential distribution of interspike intervals for the frog sacculus.
baseline activity in mammalian ANFs suggests that As mentioned above, sensory information from a
their activity in the absence of sound is driven by single IHC in the mammalian cochlea diverges onto
independent random events (i.e., it is a Poisson pro- multiple ANFs, which can be regarded as a set of
cess), except during a neural refractory period parallel channels that relay information from a single
following each spike (Johnson, D. H. and Kiang, N. IHC to the brain. Because each IHC has a slightly
Y. S., 1976) rather than by an oscillation in the pre- or different CF, and a slightly different phase relation-
postsynaptic membrane potential or by any of the ship to the sound, restricting each ANF’s input to a
other well-known mechanisms that underlie bursting single IHC preserves these two important features of
or rhythmic firing in neurons. The independent the cochlear tonotopy. Having multiple outputs from
events may be the brief openings of individual pre- each IHC with different sensitivities (range fractio-
synaptic calcium channels, in which case the afferent nation) may be important for maintaining a high
synapses onto ANFs with higher SRs could have signal-to-noise ratio for faint sounds in some ANFs
more calcium channels or a higher probability of while avoiding saturation by loud sounds in other
transmitter release in response to the opening of a ANFs. Even if some of the ANFs have identical
single calcium channel. Another possibility to explain response properties, maintaining redundant informa-
the correlation between an ANF’s sensitivity and its tion channels can be important for increasing the
SR is that the voltage dependence of the calcium signal-to-noise ratio. Two important sources of
channels at high-SR synapses is shifted such that noise are uncertainty in when (or whether) a spike
the channel’s resting open probability is increased, will be generated in a given ANF and variation in the
which would also move the resting potential to a conduction time of ANF spikes to the brain. This
steeper place on the channel’s sigmoidal activation noise can, in principle, be reduced by combining
curve. In all of these cases, high SR would be asso- information transmitted over separate channels that
ciated with greater sensitivity to sound. Since a single have independent noise sources. Each ANF postsy-
hair cell can make synapses onto several ANFs that naptic to a given IHC shares a common input signal
have widely varying SRs, these differences must be (the hair cell’s membrane potential), but the subse-
specific to individual synapses. quent steps of synaptic transmission, spike generation
If the Poisson events that trigger exocytosis are and conduction to the brain are carried out indepen-
the openings of individual calcium channels, then a dently in each AFN, and therefore are subject to
simple calculation shows that the probability of exo- independent noise. The innervation pattern of the
cytosis associated with each channel opening is likely mammalian cochlea thus appears to be optimized to
to be very small. Calcium channels in hair cells have provide the spectral and temporal information
mean open times of <1 ms and a steady-state open needed for complex computations involved in speech
probability of 10% at the resting potential, recognition and generating a map of auditory space.

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
372 Afferent Synaptic Mechanisms

The frog sacculus has a very different structure of hair cells, including IHCs of the mammalian
and pattern of innervation (reviewed in Smotherman, cochlea, are expected to be nearly isopotential even
M. S. and Narins, P. M., 2000). It is (in addition to its to the tips of the stereocilia (Roberts, W. M. et al.,
function as a vestibular organ) exquisitely sensitive to 1990). This means that the electrical signal for synap-
both airborne sound and substrateborne vibration tic transmission is the same at each of the afferent
(Koyama, H. et al., 1982). The frog saccular hair synaptic sites, and implies that the different properties
cells reside in a roughly circular patch of sensory of the afferent synapses on a single hair cell must arise
epithelium beneath the otolithic membrane, and all downstream of the electrical signal. Similarly, the
are attached to it through their kinocilia. There is no membrane capacitance is expected to attenuate high-
evidence of tonotopic organization. Instead, the hair frequency voltage oscillations to the same degree at all
cells fall into only a few anatomical classes, suggest- AZs in a given hair cell, leaving only a steady depolar-
ing that many may be functionally equivalent to each ization in response to high-frequency sounds (Section
other (Edmonds, B. et al., 2000). Unlike the mamma- 3.22.6).
lian cochlea, each saccular afferent neuron branches The hair cell’s large radius of membrane curva-
to contact many hair cells (Lewis, E. R. et al., 1982). ture compared to the diameter of a single synaptic
Summation of synaptic inputs from multiple hair AZ (several micrometers vs. hundreds of nanometers)
cells is expected to increase the sensitivity and, per- simplifies the understanding and modeling of afferent
haps more important, the signal-to-noise ratio of synaptic transmission because for some purposes the
ANFs from the frog sacculus, which is expected to membrane can be regarded as a planar boundary
contribute to the organ’s extreme sensitivity to sound separating two infinite spaces (Neher, E. and
and vibration. Augustine, G. J., 1992; Roberts, W. M., 1994).
Clearly, the hair cell’s 1 pl volume is less than
infinite and cannot serve as an infinite sink for cal-
3.22.3 Hair Cells cium entering at synapses, but as long as calcium
extrusion can keep pace with calcium entry, the
Hair cells are part of the tight epithelium that sepa- details of how and where calcium is pumped back
rates endolymph from perilymph. Each hair cell is into the extracellular space are expected to be less
sealed to its neighbors by tight junctions that form an important than in the confined space of a typical
extracellular diffusion barrier across the epithelium presynaptic bouton.
and divides the cell membrane into an apical domain
that is primarily involved in mechanosensory trans-
duction and a basolateral domain where the synapses
3.22.3.2 The Synaptic Body and Synaptic
reside (Figure 1(b)). Hair cells are both pre- and
Vesicle Pools
postsynaptic, and have other neuronal properties,
including voltage- and calcium-gated ion channels Afferent synapses on all hair cells and related elec-
that, in concert with the transduction current and the troreceptors are distinguished by a prominent
membrane’s electrical capacitance, determine how presynaptic structure, the synaptic body (SB). One
the membrane potential changes with time. The or more SBs are found at each site of synaptic contact,
hair cell’s compact shape has important consequences in close association with the presynaptic plasma
for both the electrical signal and the ensuing calcium membrane (Figures 3 and 4). Numerous functions
signal that controls transmitter release at afferent have been proposed for the SB (Sjostrand, F. S.,
synapses. 1958; Bunt, A. H., 1971; Gray, E. G. and Pease, H.
L., 1971; Osborne, M. P. and Thornhill, R. A., 1972;
Wagner, H.-J., 1978; King, T. S. and Dougherty, W. J.,
3.22.3.1 Hair Cell Shape in Relation to
1982; Schmitz, F. et al., 2000; Lenzi, D. and von
Synaptic Signaling
Gersdorff, H., 2001; Parsons, T. D. and Sterling, P.,
The hair cell body has a diameter-to-length ratio that 2003), but none proved.
is too large to allow significant spatial variation in the The hair cell contains a large population of
membrane potential, except in OHCs of the mamma- 40 nm, clear-core vesicles that are presumed to
lian cochlea, where extensive internal membranes contain the neurotransmitter, glutamate. These SVs
lining the plasmalemma might impede intracellular have been partitioned into several conceptual pools,
current flow (Nakagawa, T. et al., 2006). Other types defined either structurally with EM or functionally

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
Afferent Synaptic Mechanisms 373

(a) (b) (c)

200 nm
(d) (e)

200 nm

Figure 3 Afferent synapse ultrastructure. All panels are electron micrographs of afferent synapses in frog saccular hair
cells. (a) Each site of synaptic contact between a hair cell (top) and afferent nerve ending (bottom) typically contains one
synaptic body, which is darkly stained and surrounded by 40 nm-diameter vesicles. A second synaptic body (SB) is faintly
visible to the left. The region of close apposition between pre- and postsynaptic membranes is marked by a fuzzy appearance
that extends 100–400 nm beyond the edge of the synaptic body, but the presumed region where synaptic vesicles release
their contents (the active zone) is restricted to the area beneath the SB and perhaps 100 nm beyond. The presynaptic
membrane at the active zone contains several densities that anchor the SB. (b) The SB and its attached vesicles remain
anchored to the presynaptic active zone after the cell is broken open in extracellular saline (contains 1.8 mM calcium) and the
cytoplasm washed away prior to fixation. (c) A freeze-fracture electron micrograph shows rows of particles in the presynaptic
membrane that may align with the presynaptic densities and correspond to ion channels. The membrane dimples around the
periphery of the active zone are most likely the openings of membrane tubules rather than sites of vesicle fusion, which are
expected to occur closer to the ion channels at the active zone. (d, e) Filamentous links tether the SB to the cytomatrix at the
active zone (short arrows) and synaptic vesicles to the SB (long arrows). Synapses that had been inhibited before and during
fixation (d) had a high density of docked vesicles in the space between the SB and the plasmalemma (arrowhead points to
one). Synapses that had been stimulated before and during fixation (e) had few docked vesicles and numerous cisternal
structures in the cytoplasm. (c) Reproduced with permission from Roberts W. M. 1994. Localization of calcium signals by a
mobile calcium buffer in frog saccular hair cells. J. Neurosci. 14, 3246–3262. Copyright by the Society for Neuroscience. (d,e)
Reprinted from Lenzi, D., Crum, J., Ellisman, M .H., and Roberts, W. M. 2002. Depolarization redistributes synaptic membrane
and creates a gradient of vesicles on the synaptic body at a ribbon synapse. Neuron 36, 649–659, Copyright (2002) with
permission from Elsevier.

with electrophysiological assays of SV fusion and At both types, maximal activation of the presynaptic
neurotransmitter release. calcium current exhausts a small SV pool, defined as
Experiments on a wide variety of chemical the RRP, within a few milliseconds. At conventional
synapses in many animals have shown that SVs can synapses, the RRP comprises 1–2% of the total vesi-
generally be assigned to one of three groupings: the cle population, but considering the hair cell
readily releasable pool (RRP), the recycling pool, and cytoplasm as an unrestricted volume for SV move-
the reserve pool (reviewed in Rizzoli, S. O. and Betz, ment containing hundreds of thousands of SVs
W. J., 2005). Parallel definitions of these pools at hair (Lenzi, D. et al., 1999; Holt, M. et al., 2004), the RRP
cell afferent synapses are useful to facilitate compar- in hair cells comprises a miniscule fraction of the
isons with more conventional chemical synapses, total (Moser, T. and Beutner, D., 2000; Schnee M.
although this is hampered somewhat by their E. et al., 2005; Rutherford, M. A. and Roberts, W. M.,
different anatomical and functional characteristics. 2006).

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
374 Afferent Synaptic Mechanisms

(a) Outlying vesicles ( ) concentrate at synapses At conventional synapses, the RRP is generally
Synaptic Nonsynaptic
thought to correspond to the docked vesicles, those
seen in EM to be in contact with the plasmalemma at
the AZ. The corresponding pool of morphologically
docked vesicles at ribbon synapses consists mostly of
vesicles that occupy the space between the SB and
the plasmalemma at each AZ (Figures 3 and 4).
Differences in the precise criteria used for anatomical
classification of vesicles can cause estimates of the
size of the docked vesicle pool to differ by a factor of
two, even in careful quantitative studies.
5.2% of dense-packed 0.4% of dense-packed There is ambiguity about how to classify vesicles
(b) (c) that are tethered to the SB but do not contact the
plasma membrane. Direct evidence that these vesi-
cles even participate in transmitter release under
normal conditions is scant (Siegel, J. H. and
Brownell, W. E., 1986), but assuming that they do,
are they part of the RRP, the recycling pool, or some
other pool? One reason to think that even though
they reside far from the plasmalemma they might still
(d) (e) be in the RRP is that they may be able to fuse with
each other, forming a chain that connects to the cell
surface (reviewed in Parsons, T. D. and Sterling, P.,
2003; Edmonds, B. W. et al., 2004). Figure 4(c) shows
what may be a chain of fused vesicles. If the morpho-
logical criterion that has been used to determine
whether a vesicle is docked to the plasmalemma is
applied to vesicles on the SB, then many of these
vesicles are morphologically docked to each other,
and thus could conceivably be ready to fuse together,
Figure 4 Tomographic reconstructions of afferent although the calcium signal is expected to be highly
synapses in frog saccular hair cells. (a) Tethered vesicles attenuated more than a few vesicle diameters from the
(yellow) surround the SB (blue), which lies close to the
calcium channels on the cell surface (Roberts, W. M.,
plasmalemma (red) at the Active zone (AZ). Other (outlying)
synaptic vesicles (green) are also highly concentrated within 1994). Alternatively, the type of structures shown in
a few hundred nanometers of the SB, and populate the rest Figure 4(c), which were observed after extended
of the cytoplasm at much lower density. (b) An inhibited depolarization in high potassium (Lenzi, D. et al.,
synapse shows mostly spherical vesicles around the SB. (c) 2002), could reflect membrane retrieval.
A stimulated synapses has numerous irregular membrane
The available physiological evidence indicates
structures near the SB that could be the result of compound
vesicle fusion and/or membrane retrieval. (d, e) that the RRP in the frog sacculus and mouse cochlea
Tomographic reconstructions give a clear view of the more is much smaller than the total pool of SB-associated
extensive cisternal structures (magenta) at stimulated vesicles, and is close to the number of vesicles docked
synapses (e) compared to inhibited synapses (d). Coated at the plasmalemma (Khimich, D. et al., 2005;
vesicles (orange) and mitochondria (dark blue) are also
Rutherford, M. A. and Roberts, W. M., 2006), which
shown. Other colors are the same as in (a). (a) Reproduced
with permission from Lenzi, D., Runyeon, J. W., Crum, J., supports the more conventional hypothesis that only
Ellisman, M. H., and Roberts, W. M. 1999. Synaptic vesicle vesicles docked at the plasmalemma are part of the
populations in saccular hair cells reconstructed by electron RRP, and that other SB-associated vesicles are part of
tomography J. Neurosci. 19, 119–132; Copyright by the the recycling vesicle pool that replenishes the RRP.
Society for Neuroscience. (d,e) Reprinted from Lenzi, D.,
Physiological correlates of these anatomical defini-
Crum, J., Ellisman, M. H., and Roberts, W. M. 2002
Depolarization redistributes synaptic membrane and tions are discussed in more detail in the following
creates a gradient of vesicles on the synaptic body at a sections of this chapter.
ribbon synapse. Neuron 36, 649–659; Copyright (2002) with Tens of thousands of clear-core vesicles are found
permission from Elsevier. in the hair cell cytoplasm within several hundred

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
Afferent Synaptic Mechanisms 375

nanometers of the cell surface, and hundreds of thou- with fast binding kinetics is required to interfere with
sands more reside in the vast cell interior. They exocytosis from the RRP, whereas refilling of the RRP
occur at a low average density in frog saccular hair can be inhibited by low concentrations of slower buffer
cells (0.4% of dense packed ¼ 100 vesicles per (Moser, T. and Beutner, D., 2000). It is concluded that
mm3), but are much more concentrated within although both processes are regulated by calcium,
700 nm of each synapse (5–10% of dense packed ¼ refilling of the RRP involves signaling over a longer
1200–2400 vesicles per mm3; Lenzi, D. et al., 1999; distance, and is therefore more susceptible to disrup-
2002). The contribution of these vesicles to synaptic tion by calcium buffers. To understand the interaction
transmission under normal physiological conditions between calcium ions and calcium buffers that can
is unclear. At conventional synapses, there is a phy- diffuse in cytoplasm one must think about the rates of
siologically distinct (but sometimes not anatomically binding and diffusion, not just equilibrium binding of
distinct; Rizzoli, S. O. and Betz, W. J., 2005) recycling calcium to the buffer. A brief introduction to the essen-
pool of SVs that are reused preferentially by local tial concepts needed to understand the nonequilibrium
endocytosis (Pyle, J. L. et al., 2000; Stevens, C. F. and interaction between a mobile calcium buffer and the
Williams, J. H., 2000), whereas vesicles in the reserve calcium ions that enter through a calcium channel and
pool enter the exocytic pathway only under strong diffuse into a large cytoplasmic volume is given below.
stimulus conditions that also cause bulk endocytosis High concentrations of calcium-binding proteins
and vesicle reconstitution through an endosomal can limit the spread of calcium signals within a cell
compartment. It remains unknown whether hair cell (Roberts, W. M., 1994). This may be particularly
afferent synapses have distinct recycling and reserve important in hair cells, because they use calcium sig-
SV pools. nals for several distinct cellular processes, including
The AZ at most brain synapses contains only a few feedback inhibition of the transduction current, which
docked vesicles, and a single presynaptic action poten- involves calcium entry and action at the stereociliary
tial seldom evokes exocytosis of neurotransmitter tips (see Chapter Hair Cell Transduction and
from more than one of them, with frequent failures
Adaptation: Physiology and Molecular Mechanisms),
(Dobrunz, L. E. and Stevens, C. F., 1997; for a review
and efferent synaptic transmission see Chapter Efferent
see Sudhof, T. C., 2000). In contrast, the docked vesi-
System at postsynaptic sites, which intermingle with
cle pool is considerably larger at AZs in hair cells of
afferent presynaptic sites on the hair cell’s basolateral
the turtle papilla (mean of 30; Schnee, M. E. et al.,
surface. In addition to localized transmembrane cal-
2005), mouse cochlea (16–30, IHC; Khimich, D. et al.,
cium signals, hair cells also exhibit localized calcium-
2005), and the frog sacculus (mean of 43; Lenzi, D.
induced calcium release from intracellular stores near
et al., 2002). Furthermore, both pre- and postsynaptic
sites of efferent contact (Lioudyno, M. et al., 2004).
assays of SV exocytosis (see Sections 3.22.5–3.22.7)
The equilibrium interaction between calcium and a
have indicated that more than one SV can fuse in
molecule that has a single calcium-binding site is char-
response to a single, brief presynaptic stimulus
acterized by a single number, the dissociation constant,
(reviewed in Fuchs, P. A., 2005). The stimulus to
open voltage-gated calcium channels (VGCCs) in KD (units: molar, M), or its reciprocal, KA (M1):
hair cells is not an all-or-none action potential that ½Ca½B
lasts on the order of 1 ms, rather, they respond to KD ¼
½CaB
depolarizations of varying duration and intensity
with graded release of neurotransmitter. Thus, while where [Ca] and [B] are the concentrations (M) of free
some generalizations can be made across all synapse Ca and free buffer, and [CaB] is the concentration of
types, there may be important differences between SV calcium bound to the buffer. Free means that they are
pools at conventional synapses and hair cell afferent able to interact with each other (i.e., not already
synapses. bound to something else, hidden inside an organelle,
etc.); whether or not they are free to diffuse in the
cytoplasm is a separate issue that is also important.
3.22.3.3 Calcium Buffering in Hair Cells
Time does not appear in the units of these variables.
Exogenous calcium buffer can be introduced into a hair It is assumed that enough time has elapsed for equi-
cell through a micropipette recording electrode, and librium to be attained.
has been shown to influence synaptic transmission. The equilibrium equation does not apply near an
Interestingly, a high concentration of calcium buffer open calcium channel because equilibrium is never

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
376 Afferent Synaptic Mechanisms

reached while the calcium current persists (which the binding kinetics are important and the buffer’s
requires an extrusion mechanism to keep the cell effects can be complex. The diffusion coefficient for
from filling up with calcium). The existence of a the buffer in cytoplasm is also important because
calcium current is prima facie evidence of disequili- slow diffusion reduces the rate at which calcium-
brium, at least locally around the mouth of the free buffer is replenished near the open channel,
channel. We will not specifically consider the cal- which reduces the local [B], and in turn locally
cium pump here, except to note that the maximum increases both  and . At hair cell afferent synapses,
flux of calcium ions through a single pump exocytosis of RRP vesicles falls into the first cate-
(103 s1) is roughly one ten thousandth the flux gory, where cytoplasmic buffers are irrelevant, but
through an open calcium channel (107 s1), so that some other important calcium-dependent steps in the
the local calcium gradient around a single pump is SV cycle appear to fall into the complex category,
miniscule compared to the gradient near the mouth where buffer concentration, binding kinetics, and
of a single calcium channel. Therefore, it makes sense diffusion coefficient all come into play.
to consider the calcium extrusion process to be dis- To assign approximate values to  and  one
tributed over a large region of cell surface, and ignore needs to know the concentration and kon for the
its effect on the localized region of extremely high predominant diffusionally mobile calcium buffer(s)
free calcium near an open calcium channel. in the cell. Many types of hair cells are intensely
Two parameters, kon (M1 s1) and koff (s1), are labeled by fluorescent antibodies against one or
needed to describe the kinetics of calcium binding to more of the diffusible calcium-binding proteins
a single-site buffer. kon[B] and koff are, respectively, calbindin-D28k, calretinin, parvalbumin- , and par-
the rate constants for calcium binding to and unbind- valbumin- (also known as oncomodulin or PV-3).
ing from the buffer. The three parameters, kon, koff, For example, some hair cells in the frog sacculus
and KD are related by KD ¼ (koff/kon), so any two contain calretinin at a concentration above 1 mM,
specify the third. which provides more than 5 mM calcium-binding
A calcium ion is initially free as it enters the cell sites (Edmonds, B. et al., 2000). Other frog saccular
through a calcium channel. The mean time required hair cells contain predominantly parvalbumin , at
to encounter and bind to a buffer molecule is the extraordinary concentration of 4 mM (Heller, S.
et al., 2002), which amounts to 8 mM calcium-binding
1/(kon[B]), which for the buffering conditions
sites. OHCs in the mammalian cochlea contain different
encountered in many hair cells ([B] >> KD) is close
calcium-binding proteins than are found in the IHCs,
to the equilibration time constant () in response to a
and appear to express them at a tenfold higher concen-
step change in free calcium:
tration (calbindin-D28k and parvalbumin- equivalent
1 to 5 mM calcium-binding sites; Hackney, C. M. et al.,

kon ð½B þ KD Þ 2005). These extreme concentrations of calcium buf-
fer are certain to have important consequences for
During this time, a free calcium ion diffuses a many calcium signals in the cell. It is also worth
characteristic distance of remembering that some calcium-binding proteins
pffiffiffiffiffiffiffiffiffiffi may serve both as buffers and as calcium sensors
¼ DCa
that influence exocytosis or participate in other
where DCa is the diffusion coefficient for free calcium steps in the SV cycle.
in cytoplasm (Neher, E. and Augustine, J. G, 1992; For typical values of KD ¼ 1 mM and
Roberts, W. M., 1993). The parameters  and  are, kon ¼ 107 M1 s1, the buffer space constant and
respectively, the time and space constants for calcium time constant in a cell that has a moderate calcium
binding to the buffer in the vicinity of an open buffer concentration of [B] ¼ 100 mM are  ¼ 450 nm
calcium channel. and  ¼ 1 ms. Increasing the buffer concentration
If the distance between the presynaptic calcium 100-fold to 10 mM decreases  to 44 nm and  to
channels and the calcium sensors that trigger exocy- 10 ms. From this calculation, one can see that an
tosis is substantially less than  then the buffer has extremely high concentration of a calcium buffer
little effect. If the distance is substantially greater can reduce the buffering space constant from
than  then the buffer has time to come into equili- approximately the diameter of the entire AZ
brium with global free calcium and binding kinetics (Figure 3(c)) to approximately the diameter of a
are not important. If the distance is similar to  then single SV, thereby minimizing the overlap of calcium

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
Afferent Synaptic Mechanisms 377

domains around each open calcium channel, particu- 3.22.4 Ribbon Synapses – A Small
larly at barely perceptible sound intensities where but Diverse Class
only a few calcium channels are open at each afferent
synapse. The close association between calcium The SBs of hair cells are homologous to the presy-
channels and AZs (Lewis, R. S. and Hudspeth, A. J., naptic ribbons found in rods, cones, and bipolar cells
1983; Art, J. J. and Fettiplace, R., 1987; Roberts, W. M. in the retina, and pinealocytes in the pineal body.
et al., 1990; Issa, N. P. and Hudspeth, A. J., 1994; Like hair cells, these cells usually do not make action
Martinez-Dunst, C. et al., 1997; Rodriguez- potentials. Because of their anatomical, physiological,
Contreras, A. and Yamoah, E. N., 2001; Zenisek, D. and molecular similarities, we include the afferent
et al., 2003; Brandt, A. et al., 2005) means that steep synapses of hair cells in the class of ribbon synapses,
spatial calcium gradients and the nonequilibrium but one should also keep in mind the differences,
calcium binding to buffer are important at afferent including, for example, the slow-time scale of the
synapses. visual system compared to the auditory system.
At distances on the order of  from an open Ribbon synapses are understood to be specialized
channel, the buffer serves as a calcium sink, and can for sustained transmitter release. For example, verte-
greatly reduce the free calcium concentration brate photoreceptors release neurotransmitter
(Roberts, W. M., 1994). At larger distances, where continuously in the dark (Rea, R. et al., 2004) and
free calcium falls below the buffer’s KD, the buffer the hair cell afferent synapse is active in the absence
becomes a calcium source as it dumps its bound of an applied stimulus (Siegel, J. H., 1992). However,
calcium, although the large volume available to the hair cell is also capable of releasing transmitter
receive this calcium means that the calcium rise phasically (e.g., EPSPs that are correlated with the
associated with unbinding from the buffer is orders periodicity of the stimulus for afferent fibers at fre-
of magnitude smaller than the calcium concentration quencies up to a few kilohertz; Figures 1(c) and 2(c),
near the source. Because the buffer molecules are see also Section 3.22.6). The details of synaptic phy-
large compared to a calcium ion, bound calcium siology that vary among ribbon synapses from
different tissues and cell types are likely to be essen-
diffuses more slowly than free calcium. The net
tial for their function (for a review, see Sterling, P.
result is that the buffer can greatly increase total
and Matthews, G., 2005). Nevertheless, the first
calcium near an open calcium channel, and can
synapse in the pathways for vision, hearing, balance,
slightly increase free calcium at certain distances
and lateral line senses in all vertebrate animals is a
from the channel, but its main effect is to greatly
ribbon synapse.
reduce the extremely high concentration of free cal-
cium that accumulates near the channel (i.e., at
distances on the order of ).
More detailed discussions of these and other prop- 3.22.4.1 The Synaptic Body (Ribbon) and
erties of mobile calcium buffers are in the literature Its Vesicles
(e.g., Stern, M. D., 1992; Roberts, W. M., 1993; 1994; The term ribbon originates from the flat, almost 2D
Dargan, S. L. and Parker, I., 2003). One interesting morphology of the presynaptic structures observed in
result is that the average distance that a calcium ion transmission EMs of retinal photoreceptors and bipo-
travels while remaining bound to the buffer is >1 mm, lar cells. Serial reconstructions of gluteraldehyde-
which is large compared to the AZ diameter. Thus, fixed hair cells from many preparations, including
the mobile buffer gives a calcium ion only one brief mammalian cochlea, have shown that the SB in hair
chance to exert its effect. Either it binds to its target, cells has a more spherical or ellipsoid shape. IHCs
such as the calcium sensor that triggers exocytosis, or rely exclusively on ribbon synapses for afferent trans-
it is captured by the mobile buffer and is carried mission. The role of SBs in OHCs is more uncertain
away. If the buffer has reasonably fast binding as they (the SBs) are far fewer and often seen only
kinetics (kon ¼ 107 M1 s1) and provides 5–10 mM during development (Smith, C. A. and Sjostrand, F. S.,
of mobile binding sites for Ca (Edmonds, B. et al., 1961; Siegel J. H. and Brownell W. E., 1981;
2000; Heller, S. et al., 2002), then the average time Sobkowicz, H. M. et al., 1986).
available for a calcium ion to encounter and bind to Although SBs are heavily stained by osmium,
its target (i.e., the binding time constant, ) is only which typically labels lipid membranes, they are
10–20 ms. thought to be composed primarily of protein

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
378 Afferent Synaptic Mechanisms

(Krstic, R., 1976). The osmiophilic nature of the SB is forms 50 nm-long rods in vitro (Yamazaki, H. et al.,
well-known because OsO4 is a common EM fixative 1995), or possibly one or more of the many other
and is electron-dense. Perhaps more relevant to func- KIF3 isoforms that may cross react with the antibody.
tion, the ribbon has been reported to bind both SBs and ribbons can be regarded as elaborations of
calcium and phosphate (Schmitz, F. and the cytomatrix at the active zone (CAZ) seen at all
Drenckhahn, D., 1991), although there is no evidence synapses (Zhai, R. G. and Bellen, H. J., 2004), but in
that it binds or releases these ions in the process of EM the SB appears as a separate structure, connected
synaptic transmission. to the CAZ by filamentous links (Usukura, J. and
SBs at AZs are always found in association with Yamada, E., 1987; Lenzi, D. et al., 2002). The tight
40 nm-diameter clear-cored vesicles, which are association between the SVs and the SB, and between
attached to it by thin filaments (Usukura, J. and the SB and the presynaptic plasma membrane is
Yamada, E., 1987) and often cover its surface nearly evident in cells that were broken open prior to fixa-
completely (Figures 3 and 4; Lenzi, D. et al., 2002). tion (Figure 3(b)). If the binding were weak or
SBs can sometimes be found in the cell interior, disrupted by exposure to the high-calcium concen-
either singly or in clusters, where they are also cov- tration of extracellular saline, these structures would
ered in vesicles (Miller, M. R. and Beck, J., 1987). The not have remained behind while the rest of the cyto-
molecular identity of the filaments is not known, plasm was washed away.
except that they are thought not to contain synapsins
I or II, which form filaments linking SVs to cytoske-
3.22.4.2 Molecular Specializations of
leton at other synapses (Hirokawa, N. et al., 1989), but
Ribbon Synapses
are absent from ribbon synapses (von Kriegstein, K.
et al., 1999). Based on antibody labeling, the filaments Understanding of the molecular composition of hair
could contain the KIF3A subunit of the kinesin 2 cell afferent synapses has begun (Figure 5), but is far
motor complex (Muresan, V. et al., 1999), which from complete (Wagner, H.-J., 1997). A protein, Ribeye

(a) (b) Ribbon


RIBEYE/CtBP2
Hair cell CtBP1
KIF3A
Piccolo
RIM1

CAZ
Ribbon Bassoon
RIM2
VGCCs
Munc13-1
Glast BKs
CAST1

Observed
in hair cells
AMPARs
Support Inferred from
cell Afferent bouton retinal ribbons

Figure 5 Molecular specializations of the hair cell ribbon synapse. (a) Localization of proteins associated with the synaptic
ribbon and plasma membrane density of the cytomatrix at the active zone (CAZ), labeled in (b). Voltage-gated calcium channels
(VGCCs) are clustered underneath the ribbon and juxtaposed to postsynaptic -amino-3-hydroxy-5-methylisoxazole-4-
propionic acid (AMPA)-type glutamate receptors. In some hair cells, presynaptic calcium-sensitive voltage-gated (BK)
potassium channels intermingle with the VGCCs. Glutamate transporters in supporting cells clear neurotransmitter from the
cleft. Bassoon anchors the ribbon to the Cytomatrix at the active zone (CAZ) plasma membrane density via its interaction with
the RIBEYE. Bassoon, RIBEYE/CtBP2, CtBP1, and Piccolo have been observed at hair cell afferent synapses. The existence of
the other proteins at the hair cell’s ribbon synapse is by inference from observations at retinal ribbon synapses (tom Dieck, S.
et al., 2005). Reproduced from The Journal of Cell Biology, 2005, 168, 825–836 by Copyright permission of The Rockefeller
University Press; and Reprinted from Curr. Opin. Neurobiol. Vol. 13, Fuchs, P. A., Glowatzki, E., and Moser, T., The afferent
synapse and cochlear hair cells, 452–458, Copyright 2003, with permission from Elsevier.

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
Afferent Synaptic Mechanisms 379

(Schmitz, F. et al., 2000), has been identified that Besides Ribeye, only one other protein, the B16
appears to be a component of all synaptic ribbons, antigen, has been proposed to be a unique component
and is expressed nowhere else. This finding justifies of all ribbon synapses. This protein is known only as an
exclusion of T-bar synapses of arthropods (e.g., antigenic epitope (amino sequence: DTYQHPPKDS;
Drosophila) from the class of ribbon synapses because Balkema, G. W., 1991; Nguyen, T. H. and Balkema, G.
the Drosophila genome contains no DNA sequences that W., 1999). A search of the human genome and
code for a homolog of the unique exon 1 of the Ribeye expressed sequence tag (EST) databases shows that
protein (Bradford, Y. and Roberts, W. M., unpublished only one gene, aconitase, contains a nucleotide
observations). Ribeye is one of two known protein sequence that codes for this exact amino acid sequence.
products of the CtBP2 gene. The other product, Aconitase is a transcription regulator and a regulator of
CtBP2, is a transcription regulator (reviewed in iron metabolism, but reportedly is not present in
Chinnadurai, G., 2002). A recent report suggests that synaptic ribbons (Nguyen, T. H. and Balkema, G. W.,
the CtBP1 protein, which is the product of a closely 1999). Human CtBP2 and Ribeye contain a sequence
related gene and also a known transcription regulator, (TYxxPP) near the C-terminus that is part of the B16
is a presynaptic component of both hair cell (Moser, T., antigen and thus might bind the B16 antibody, but
personal communication) and retinal (tom Dieck, S. synaptic ribbons in fish reportedly also bind the B16
et al., 2005) ribbon synapses, as well as conventional antibody (Balkema, G. W., 1991) even though this
synapses, raising the possibility that the ancestral CtBP sequence is entirely absent from their CtBP2 and
protein had a presynaptic function in addition to being Ribeye proteins. Thus, the component of the ribbon
a transcriptional corepressor, and conceivably could that binds antibodies raised against the
link synaptic activity to gene expression. The DTYQHPPKDS peptide remains unknown.
CAZ-associated protein Bassoon was recently shown There are other proteins that may be unique to
to anchor the ribbon at the photoreceptor synapse via a ribbon synapses but are not expressed at all ribbon
physical interaction with Ribeye (tom Dieck, S. et al., synapses. The 1D calcium channel is found only in
2005). Bassoon and the related protein Piccolo, are also hair cells and is essential for normal hair cell Ca2þ
present at hair cell afferent synapses (Khimich, D. et al., current, neurotransmitter release, and for hearing
2005). Bassoon appears to interact with both Ribeye (reviewed in Moser, T. et al., 2006a). Most synapses
and CtBP1, which are both localized to ribbons even in express N, P/Q, or R-type calcium channels, but hair
the absence of functional Bassoon protein. cells express L-type calcium channels (Cav 1.3) formed
Drosophila has a single CtBP gene that produces a by this unique splice variant. This channel opens at
transcription regulator essential for embryonic unusually hyperpolarized voltages (Koschak, A. et al.,
development (Nibu, Y. et al., 1998; Poortinga, G. 2001; Brandt, A. et al., 2003; Rodriguez-Contreras, A.
et al., 1998), but it is not known whether CtBP is and Yamoah, E. N., 2003; Schnee, M. E. and Ricci, A. J.,
present at synapses in Drosophila. Similarly, the 2003), which may cause the transmitter release respon-
Ciona genome (a urochordate) has a single CtBP sible for the baseline action potential rate (i.e., SR) in
gene that lacks the coding sequence for Ribeye afferent neurons. Retinal rods express a different L-type
(Bradford, Y. and Roberts, W. M., unpublished obser- calcium channel at ribbon synapses, Cav 1.4 ( 1F;
vations). The simplest evolutionary scenario Morgans, C. W., 2001).
consistent with the data is that a gene duplication Another protein that may be unique to ribbon
event early in the vertebrate lineage, but after the synapses is complexin IV (Pabst, S. et al., 2000;
divergence of urochordates from vertebrates, gave Reim, K. et al., 2006). Based on similarity of com-
rise to paralogous genes, CtBP1 and CtBP2. The plexin I and II, and the ability to rescue synaptic
CtBP2 gene then acquired an alternate first exon, transmission in complexin I/II knockout mice, both
which became involved in aggregation of CAZ com- complexin III and complexin IV are potential cal-
ponents (Schmitz, F. et al., 2000) into large structures cium-dependent regulators of synaptic transmission.
(ribbons), with important, but largely unknown con- Complexins I and II are known to bind to the soluble
sequences for afferent synaptic transmission in the NSF (N-ethylmaleimide-sensitive factor) attachment
eye and ear. Ribeye expression became limited to the receptor (SNARE) complex that forms the core of
small class of ribbon synapses, while CtBP1 continues the vesicle fusion machinery at many synapses
to be expressed at other synapses. An additional gene (McMahon, H. T. et al., 1995; Reim, K. et al., 2001;
duplication gave rise to two CtBP2/ribeye genes in Chen, X. et al., 2002; for a review, see Sudhof, T. C.,
teleost fishes (Wan, L. et al., 2005). 2004). Complexins III and IV have both been

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
380 Afferent Synaptic Mechanisms

described at ribbon synapses in the retina, but only shape within individual cells. The dimensions of the
complexin IV was found to be unique to retinal SB are typically submicron, ranging from 200 nm
ribbon synapses. It was not reported whether com- for the shortest principle axis of the ellipsoid SBs in
plexins III or IV are present in the ear. mouse IHCs (Khimich, D. et al., 2005) to 400 nm in
Otoferlin is another protein that may be involved hair cells of frog and goldfish sacculus (Gleisner, L.
in synaptic transmission by hair cells. This protein et al., 1973; Hama, K. and Saito, K., 1977; Jacobs, R. A.
was discovered by genetic mapping of familial deaf- and Hudspeth, A. J., 1990; Lenzi, D. et al., 1999).
ness (Yasunaga, S. et al., 1999), and is essential for Estimates of the number of ribbon-containing
hearing. The otoferlin gene is large and complex, synapses per cell ranged from 10 in mouse cochlear
having 48 exons and many splice variants (Yasunaga, IHCs (Khimich, D. et al., 2005) to 85 in the high-
S. et al., 2000). It is similar to the FER-1 protein of frequency hair cells of the turtle papilla (Sneary, M. G.,
Caenorhabditis elegans, which has been implicated in 1988).
vesicle membrane fusion (Achanzar, W. E. and Several studies have provided clues regarding the
Ward, S., 1997). Otoferlin is a candidate to substitute function of SBs in hair cells by demonstrating that the
for synaptotagmin as the calcium sensor for transmit- size and number of synaptic ribbons, as well as the
ter release (Yasunaga, S. et al., 2000). numbers of associated calcium channels and SVs, is
Several other presynaptic proteins, including correlated with the position of the hair cell on the
synapsins and synaptotagmins, are notable by their tonotopic axis of the organ. In the turtle papilla, high-
absence from ribbon synapses (Safieddine S. frequency cells have three times more SBs, higher
and Wenthold R. J., 1999; von Kriegstein, K. et al., vesicle densities near SBs, more calcium channels,
1999). Syntaxin 1, a component of the core SNARE and larger calcium currents than do lower-frequency
complex is also missing, replaced by syntaxin 3 cells (Sneary, M. G., 1988; Art, J. J. et al., 1995; Schnee,
(Morgans, C. W. et al., 1996). M. E. et al., 2005). Upon depolarization, the high-
frequency cells released more vesicles and with faster
kinetics. A model of SV trafficking that incorporated
3.22.4.3 Anatomy of the Synaptic Body:
all of these data predicted that these differences
Heterogeneity of Structure
result in transmission being optimized to the CF of
SBs come in a bewildering variety of shapes and sizes. the hair cell (Schnee, M. E. et al., 2005). In the papilla
Most synapses have one SB at the synaptic complex, of the chick, the diameter of the SB ranged from 120
others have two or three (Liberman, M. C. et al., 1990). to 210 nm with increasing CF from apical to more
In addition to the typically spherical or ellipsoidal basal positions along the tonotopic axis (Martinez-
shapes of SBs in hair cells from mice, turtles, frogs, Dunst, C. et al., 1997). These two studies used differ-
and chicks (Sneary, M. G., 1988; Martinez-Dunst, C. ent experimental approaches to reach a similar
et al., 1997; Lenzi, D. et al., 1999; Khimich, D. et al., conclusion; that the morphology of the SB has influ-
2005), SBs of different geometries and densities have ence over the numbers of SVs that are maintained at
been observed in EM (Giannessi, F. and Ruffoli, R., well-defined positions with respect to the source of
1996). In many types of hair cells, the SB appears as a calcium influx. SVs close to the center of the AZ may
uniformly dense blob, with no internal structure but be more responsive to calcium influx if the channels
it also can have a laminar appearance, as in some are clustered there (Furukawa, T. and Matsuura, S.,
cochlear IHCs. In old world monkeys the SBs of 1978; Furukawa, T. et al., 1982; reviewed for the case
IHCs nearly always appear ring shaped (Bodian, D., of CNS synapses in Schneggenburger, R. and Neher,
1980). It is unclear how much of this variability is an E., 2005).
artifact of fixation and staining. For example, in the A study in the mammalian cochlea has reported
frog retina, synaptic ribbons of a lamellar nature were presynaptic morphological correlates of ANF SR and
observed with conventional transmission electron threshold (Merchan-Perez, A. and Liberman, M. C.,
microscope (TEM) thin sections, but were not so 1996). Analysis of ultrastructure at the EM level
apparent when freeze-etching and freeze-substitu- showed that the size and complexity of SBs tended
tion methods were employed for tissue preparation to increase as SR decreased, accompanied by an
(Usukura, J. and Yamada, E., 1987). increase in the total number of SVs clustered at the
SBs vary in size and number among organs, within AZ. These data suggest that the size of the SB has
and between species, as well as among cells within a something to do with the regulation of vesicle release
particular organ. Furthermore, SBs vary in size and that determines the baseline afferent discharge rate

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
Afferent Synaptic Mechanisms 381

and sensitivity to faint sounds, but the relationship is preformed vesicles present throughout the cyto-
opposite what is expected based on the prevailing plasm, and thus may not have to rely on endocytic
view that the ribbon collects vesicles and conveys routes for replenishment of releasable SVs for even
them to the release sites. Instead, these data seem to the longest stimuli used in many experiments
suggest that a larger SB may somehow inhibit SV (Griesinger, C. B. et al., 2005). For example, a hair
fusion for weak sounds. Clearly, more work needs to cell that contains 1 million vesicles could, in princi-
be done to understand the relationship between SB ple, sustain the exocytosis of 10 vesicles per
morphology and postsynaptic properties. millisecond for 100 s, although is hard to imagine
how the hair cell could accommodate the resultant
threefold increase in surface area. Clearly, in the long
3.22.5 The Synaptic Vesicle Cycle in run endocytosis still must keep pace with exocytosis.
Hair Cells
3.22.5.2 Vesicle Mobility
The basic features of the SV cycle are similar
at conventional chemical synapses and ribbon The movement of vesicular membrane through the
synapses. Both baseline and evoked EPSCs at ribbon cytoplasm is an essential step in the SV cycle, except
synapses require presynaptic calcium influx through for the case of kiss-and-stay. At conventional
channels (Robertson, D. and Paki, B., 2002; Keen, E. synapses, the vesicle-associated phosphoproteins,
C. and Hudspeth, A. J., 2006) and are completely called synapsins, play essential roles in regulating
and reversibly blocked by antagonists of the -ami- the movement and clustering of SVs (Rosahl, T.
no-3-hydroxy-5-methylisoxazole-4-propionic acid et al., 1995). Studies on hippocampal synapses in
(AMPA)-type glutamate receptor (Glowatzki, E. synapsin knockout mice have shown that synapsins
and Fuchs, P. A., 2002; Keen, E. C. and Hudspeth, contribute to the SV cycle by maintaining the reserve
A. J., 2006; Holt, J. C. et al., 2006). The calcium- pool and the RRP at glutamatergic and GABAergic
controlled release of glutamate into the synaptic terminals, respectively (Gitler, D. et al., 2004). The
cleft is quantal, involving exocytosis from SVs that synapsins tether SVs to actin filaments in a phosphor-
fuse with the plasmalemma and are later retrieved for ylation-dependent manner (for a review, see Hilfiker, S.
reuse. et al., 1999).
Synapsins are absent at ribbon synapses, supporting
the idea that vesicle clustering and mobilization are
3.22.5.1 Endocytosis
different in these cells (von Kriegstein, K. et al., 1999).
Various routes for vesicle recycling at chemical Recent work in retinal bipolar cells utilized EM, con-
synapses have been proposed, including two contro- focal microscopy, and total internal reflection
versial types of fast, local reuse in which the microscopy to demonstrate that cytoplasmic vesicle
neurotransmitter leaves the SV through a fusion movement is rapid, random, and not altered by cal-
pore without full collapse of the SV membrane into cium influx or cytoskeletal interactions (Holt, M. et al.,
the plasmalemma. In one of these hypothetical fast 2004, reviewed in Sterling, P. and Matthews, G., 2005).
recycling routes, named kiss-and-stay, the vesicle The authors concluded that SVs 2 are highly mobile in
remains docked to the AZ where it is presumably bipolar cells (D ¼ 1.5  102 mm s1), and that random
refilled with glutamate after discharging its contents. vesicle movement (i.e., diffusion) is sufficient to
In the other, named kiss-and-run, the vesicle undocks account for the translocation of vesicles from the
and enters a local recycling pool. Slower recycling large cytoplasmic pool to release sites.
pathways involve full fusion of the SV with the If baseline spike rates in ANFs are due to fusion of
plasmalemma, followed by clathrin-dependent endo- SVs, then the diffusion coefficient reported for SVs in
cytosis of vesicles or bulk membrane retrieval and bipolar terminals is far too low to support continuous
reprocessing of SVs either directly or via an endoso- resupply of release sites at mammalian hair cell
mal route (reviewed in Sudhof, T. C., 2004). The synapses that sustain SRs of >100 Hz unless some
extent to which each of these processes contributes additional mechanism exists to concentrate vesicles
to the SV cycle under natural physiological condi- in the vicinity of the synapse. Even assuming that
tions remains unclear. diffusion is used only to transport vesicles to the SB,
Recent work in the cochlea showed that IHCs whereupon they are instantly transported to release
may be able to draw upon the vast supply of sites by some other means, and ignoring the

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
382 Afferent Synaptic Mechanisms

plasmalemma and other barriers that restrict diffu- kinesin-mediated translocation of vesicles to the plas-
sional access to the SB, the maximum steady-state malemma (Heidelberger, R. et al., 2002). This
rate at which vesicles can reach the SB by diffusion is observation led to the alternative hypothesis that SVs
F ¼ 4DRC
on the ribbon can fuse with each other (reviewed in
Parsons, T. D. and Sterling, P., 2003), creating chains
where C is the global concentration of vesicles in that extend to the plasmalemma. This mechanism can
the volume. Substituting D ¼ 1.5  102 mm2 s1, account for the large EPSCs recorded postsynaptically
F ¼ 100 s1, and R ¼ 100 nm, and solving for C yields at hair cell synapses (see Section 3.22.6.3), and implies
C ¼ 5000 vesicles mm3, which is more than 50 times that vesicles on the ribbon can be primed and fuse with
the density reported in the nonsynaptic regions of each other. However the protein Munc13-1, thought
hair cells. This calculation shows that the average to be necessary for priming (Rosenmund, C. et al.,
concentration of SVs throughout the cytoplasm is 2003), was found to cluster with calcium channels at
too low to continuously supply the SB with vesicles the plasma membrane and not on the ribbon (tom
by diffusion unless some additional process operates Dieck, S. et al., 2005), implying that the final priming
to maintain the locally high concentration of SVs steps must take place near the membrane.
within a few hundred nanometers of the SB (Lenzi,
D. et al., 1999; 2002).
Experiments like these, combined with the 3.22.5.4 Transmitter Release and
observed presence of synaptic ribbons in cell types Clearance
that support continuous exocytosis have led to the Glutamate released into the synaptic cleft binds to
view that the structural and molecular specializations postsynaptic AMPA receptors (Ruel, J. et al., 2000;
of ribbon synapses assist high rates of sustained vesicle Glowatzki, E. and Fuchs, P. A., 2002; Keen, E. C. and
supply to AZs, but crucial questions remain unan- Hudspeth, A. J., 2006), causing a depolarization of the
swered. For example, data for the refilling of the RRP postsynaptic membrane and, if the depolarization is
by vesicles on the SB are indirect; it is still unknown sufficient, one or more action potentials. N-Methyl-
whether or not vesicles actually move on the ribbon. D-aspartate (NMDA) receptor antagonist had no
effect on ANF activity, suggesting that they are
functionally absent. The pattern of spikes depends
3.22.5.3 Docking and Priming
on the temporal pattern of neurotransmitter release,
In addition to translocation to the AZ, other things the kinetics of AMPA receptor activation, stochastic
must occur before a vesicle is ready for exocytosis. spike generation, and neuronal refractoriness. Most
Two important but relatively poorly understood steps in vivo studies have focused on the firing properties of
are called docking and priming, each of which likely ANFs in response to pure tones of varying intensities.
embodies a series of associated events. Docking refers If quantal release of neurotransmitter is to accurately
to the placement of a SV in contact with the inner reflect the temporal pattern of the sound stimulus, then
leaflet of the plasma membrane whereas priming the chemical signal in the synaptic cleft must rapidly
refers to the molecular reactions, some ATP-depen- be terminated, which involves uptake of glutamate
dent, required to make a vesicle immediately ready for from the synaptic cleft by excitatory amino acid
fusion upon calcium influx. It is unclear whether a transporters (e.g., GLAST) in surrounding support-
vesicle must first be docked in order to be primed. ing cells and afferents (Furness, D. N. and Lehre, K.
For clues to the relationship between fast exocyto- P., 1997; Furness, D. N. and Lawton, D. M., 2003).
sis and docking/priming at ribbon synapses, we look to Voltage-clamp recordings from inner phalangeal
experimental evidence from retinal bipolar cells. Here cells in the cochlea of normal rats and GLAST
the ribbon is plate-like and the anatomical picture of knockout mice showed the presence and absence,
rows of vesicles tethered to both sides of the ribbon led respectively, of currents produced by these transpor-
to the model in which the vesicles move along the ters (Hakuba, N. et al., 2000). Normal glutamate
ribbon to docking sites. One hypothesis is that during a reuptake from the synaptic cleft is also thought to
100 ms strong depolarization all vesicles on the ribbon be important for the prevention of noise-induced
are depleted via exocytosis. Assuming that this exocy- damage to afferent terminals, which can be mimicked
tosis requires movement of vesicles along the ribbon to by GLAST blockers (Hakuba, N. et al., 2000;
the plasmalemma, several lines of evidence argue Rebillard, G. et al., 2003) and prevented by AMPA
against the involvement of ATP-dependent, receptor blockers.

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
Afferent Synaptic Mechanisms 383

3.22.5.5 Relating Anatomical and can be translated into numbers of SVs exocytosed,
Functional Vesicle Pools although given the uncertainties in SV diameter and
specific membrane capacitance, such calculations are
The hair cell’s compact shape has proved fortuitous for
probably only accurate to within a factor of two.
experimental measurements of synaptic exocytosis by
These functionally derived SV counts can be com-
observing the increase in electrical capacitance asso-
pared to numbers obtained with anatomical studies at
ciated with the increased cell surface that occurs when
the EM level in order to test hypotheses about the
SVs fuse with the plasmalemma (Figures 6(a)–6(d);
relationships between the structure and the function
Lindau, M. and Neher, E., 1988; reviewed in
of SV pools. The reliability of this technique in
Nouvian, R. et al., 2006). Because the surface area of accurately monitoring the release of neurotransmit-
a single SV can be estimated, capacitance changes ter has been verified by combining capacitance

(a) (c) (d)

20 10

Pool size (fF)


8
15
ΔCm (fF)

10 fF 6 τf = 140 ms
10
τs = 3 s
4
5
2
40 pA
0 2 4 6 8 10 32 34 0
200 ms
0 10 20 30 40 50 Time after pool depletion (s)
(b) Duration of stimulation (ms) (e) 450

300 10 ms

Discharge rate (spikes/s)


300 0.1 mM EGTA
5 mM EGTA 150
250 5 mM BAPTA
ΔCm (fF)

200 Endogenous buffer 0

150 450

100 300 160 ms


50
150
0
0
500 1000
0 300 600
Duration of stimulation (ms)
Time (ms)

Figure 6 Exocytosis from inner hair cells, recovery from synaptic vesicle pool depletion, and adaptation/disadaptation of
auditory nerve fiber responses. (a) Plasma membrane capacitance changes (upper) caused by depolarization-induced
synaptic vesicle exocytosis that depends on calcium influx (lower). Two 15 ms depolarizations to the peak of the IV curve for
calcium (15 mV) were separated by 100 ms. Although the calcium currents evoked by the paired pulses were the same, the
exocytic response to the first pulse was greater than the response to the second pulse (secretory depression). (b, c) The
amplitude of stimulus-evoked capacitance changes is plotted as a function of the duration of the depolarization for different
intracellular calcium-buffering conditions. (c) The first 50 ms of (b) at higher temporal resolution. With the cell’s endogenous
buffer present, the readily releasable pool (RRP) is depleted within the first 10 ms, after which exocytosis continues at a slower
rate that is sustained for at least 1 s. The slow calcium buffer (EGTA) of 5 mM abolishes the sustained release while leaving
exocytosis of the RRP intact. The fast calcium buffer (BAPTA) of 5 mM inhibits both fast exocytosis of the RRP and the slower
component of exocytosis. (d) Strong depolarizations sufficient to deplete the RRP were followed by a second identical
stimulus with varying interstimulus intervals as indicated on the x-axis. Recovery of the RRP as a function of time after pool
depletion has both rapid and slow components. (e) Auditory nerve fiber spike rate in vivo as a function of time after sound
onset. The sound is interrupted by 10 ms (upper) or 160 ms (lower). During a sound presentation, the spike rate is initially very
high followed by two phases of fast reduction in firing rate called rapid and short-term adaptation, with time courses of about
10 and 100 ms, respectively. The 10 ms interruption (upper) is not enough for recovery from rapid adaptation of the spike rate
whereas the 160 ms interruption (lower) is sufficient for nearly complete recovery of the initial spike rate. Note the depression
of baseline rate and its recovery after sound offset. Panels (a)–(d) are for the mouse: Reproduced with permission from,
Copyright (2000) National Academy of Sciences, U. S. A., Moser, T. and Beutner, D., Kinetics of exocytosis and endocytosis
at the cochlear inner hair cell afferent synapsis of the mouse. Proc. Natl. Acad. Sci. U S A 97, 883–888. Panel (e) is from the
chick: JARO, Vol. 5, 2004, Evidence that rapid vesicle replenishment of the synaptic ribbon mediates recovery from short-
term adaptation at the hair cell afferent synapse, Spassova, M. A., Avissar, M., Furman, A. C., Crumling M. A., Saunders, J. C.,
and Parsons, T. D., Copyright Springer, with kind permission of Springer Science and Buisness Media.

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
384 Afferent Synaptic Mechanisms

measurements with simultaneous detection of gluta- Exocytosis of the SSC is slow compared to the
mate release in bipolar cells of the vertebrate retina RRP (but still a fast resupply rate of hundreds of
(von Gersdorff, H. et al., 1998). vesicles per second per AZ), and may indicate the
To estimate the size of the RRP in hair cells, synap- rate at which the RRP is replenished from a recycling
tic transmission is first inhibited by hyperpolarization pool in the presence of a continuing calcium influx.
(to fill vesicle pools), then a depolarizing voltage stimu- Dialyzing the cell with the calcium buffer EGTA to
lus that maximally stimulates exocytosis is delivered, inhibit the spread of free calcium away from the AZ
and the ensuing capacitance increase is observed (Figures 6(b) and 6(c)) blocked the SSC but not the
(Figures 6(b) and 6(c); reviewed in Fuchs, P. A. et al., initial exocytic burst from the RRP (Spassova, M. A.
2003). Pairs of brief strong depolarizations that evoke et al., 2004; Moser, T. and Beutner, D., 2000), which
maximal calcium currents typically cause paired- suggests that refilling of the RRP is stimulated by
pulse depression (Figures 6(a) and 6(d)), implying long-distance calcium signaling that can be blocked
that the first stimulus partially or totally depletes by a slow calcium buffer. Exocytosis of the RRP was
the RRP, which is refilled more slowly. The size of inhibited only with the fast calcium buffer BAPTA,
the RRP can also be measured by applying strong as is expected for the case of tight spatial coupling
continuous depolarizations of various durations, between the release-ready vesicles and the calcium
which evoke continuous calcium currents from the channels (i.e., the RRP docked at the AZ). At present,
L-type calcium channels at ribbon synapses. it is unknown whether the SSC involves vesicles on
Exocytosis of the RRP is seen as an initial rapid the SB, other cytoplasmic vesicles that move to the
increase in capacitance that lasts only a few millise- AZ, or possibly even vesicles that are not located at
conds before giving way to a sustained secretory AZs. Experiments that combine capacitance mea-
component (SSC) that lasts for seconds (Figure 6(b); surements with other techniques to monitor the
Moser, T. and Beutner, D., 2000; Spassova, M. A. flow of membrane through the SV cycle will be
et al., 2004; Rutherford, M. A. and Roberts, W. M., required to answer such questions.
2006). Generally, the size of the RRP measured by
change in capacitance is in good agreement with the
size of the docked SV pool, as measured by EM, and
is smaller than the pool of vesicles on the SB. The 3.22.6 Physiology of the Hair Cell
presence of a SB and the size of the RRP are both Afferent Synapse
important for normal ANF responses (Khimich, D.
et al., 2005; see also Section 3.22.7.2). Precise encoding at the first synapse in the auditory
The density of vesicles on the SB and in the system is absolutely critical for the complete range of
cytoplasm surrounding the AZ is sensitive to the behavioral interactions with the acoustic world. To
physiological state of the hair cell before and during produce a neural representation of the acoustic
fixation (Lenzi, D. et al., 2002). In frog saccular hair environment in the CNS, the mammalian cochlea
cells, prolonged depolarization by high extracellular first decomposes complex stimuli into separate fre-
potassium reduces the density of vesicles on the SB quency components, each of which has an amplitude
from 79% to 42% of 3D dense-packed, with the and a phase. The simplest statement that can be made
greatest reduction occurring in the number of docked about sensory encoding in the mammalian cochlea is
SVs in the narrow space between the SB and the that the amplitude of the sound at a given frequency
plasmalemma. Fixation under resting conditions is encoded by the spike rate in the ANFs of corre-
yielded intermediate values (Lenzi, D. et al., 2002). sponding CF, while phase and other temporal
Whole-cell measurements of SV exocytosis support patterns (e.g., amplitude modulations) are specified
the notion that all of the docked, SB-associated vesi- by the timing of ANF spikes. Decades of work have
cles can fuse with the plasmalemma within 10 ms produced a wealth of detail about sensory encoding
(Rutherford, M. A. and Roberts, W. M., 2006). in the auditory system, but much remains to be dis-
Vesicles in the cytoplasm that are not attached to covered. The following will barely scratch the
the SB are also massively depleted by prolonged surface of what is known. For additional information,
depolarization (from 10% to 2% of 3D dense-packed the mechanisms underlying the temporal precision of
in reconstructed volumes within 800 nm perpendicu- sound coding at the IHC ribbon synapse have been
lar to the plasmalemma; Figure 4). recently reviewed (Moser, T. et al., 2006b).

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
Afferent Synaptic Mechanisms 385

3.22.6.1 Encoding of Phase frequency can be attributed to the low-pass filtering


of the AC component of the receptor potential by the
For low-frequency sinusoidal (tonal) stimuli, action
hair cell membrane (Figure 7). When intracellular
potentials in ANFs occur within a narrow time window
receptor potentials were recorded in vivo from a gui-
relative to each cycle of the sinusoid (Anderson, D. J.
nea-pig cochlear IHC, the receptor potential was
et al., 1970). The discharge of the ANF is said to be
quasisinusoidal with the periodicity of the stimulus
phase-locked to the stimulus (Rose, J. E. et al., 1967).
for low-frequency tones (Palmer, A. R. and Russell, I.
Phase locking has been observed experimentally in the J., 1986; Cheatham, M. A. and Dallos, P., 1993). The
auditory pathways of mammals, birds, reptiles, amphi- AC (periodic) component of the receptor potential
bians, and fish, and is thought to be an important began to decline at 600 Hz and was not detectable
mechanism for the temporal encoding of low-frequency above 3.5 kHz, leaving only a DC (steady) depolar-
sound. For example, phase differences between the two ization (Figures 7(a) and 7(b)). The diminution of the
ears provide important binaural cues for localizing presynaptic AC component was mirrored by a
sounds horizontally (Knudsen, E. I. and Konishi, M., decline in postsynaptic phase locking (Figure 7(c)).
1978; see Chapter Encoding of Interaural Timing for Recordings of spike trains in the saccular branch of
Binaural Hearing). the eighth nerve in frogs have shown that vibrations at
Phase locking declines as frequency increases until frequencies 50 Hz evoke a single, phase-locked spike
action potentials in the auditory nerve become nearly for each cycle of the stimulus (Koyama, H. et al., 1982;
random with respect to the stimulus period. The Christensen-Dalsgaard, J. and Narins P. M., 1993).
frequency where this transition occurs in the cochlear When expressed as spikes per second, the response
nerve is generally between 1 and 4 kHz, but it varies thus decreases in proportion to frequency below
among species and even within an individual cochlea. 50 Hz. Because for low frequencies the depolarizing
In part, the decline of phase locking with increasing phase of the hair cell receptor potential may be

(a) (b) 10.0


AC component / DC component

100
3.0
300
1.0

500
25 mV 0.3

700
0.1
0.03 0.1 0.3 1.0 3.0
900 Frequency (kHz)
(c)
1.2
1000
Synchronization index

1.0
AC
D.C. 2000 0.8
3000 0.6
4000
0.4
5000
0.2
0 10 20 30 40 50 60 70 ms 0.0
0.1 1.0 10.0
Frequency (kHz)

Figure 7 The relationship between receptor potentials in inner hair cells (IHCs) and phase locking in auditory nerve fibers.
(a) Intracellular recordings of receptor potentials in one IHC of the guinea-pig cochlea in response to loud (80 dB sound
pressure level) tones at the frequency (Hz) indicated by the right side of each trace. The receptor potential has an oscillatory
(AC) component and a steady-state (DC) component. The AC component dominates at low frequencies but is almost absent
above 3000 Hz. (b) The ratio of AC to DC component across frequency for the receptor potentials from nine different IHCs.
(c) The synchronization index (a measure of phase locking to the tonal stimulus) plotted as a function of tone frequency.
(Reprinted from Palmer, A. R. and Russell, I. J. 1986. Phase-locking in the cochlear nerve of the guinea-pig and its relation to
the receptor potential of inner hair-cells. Hear Res. 24, 1–15, with permission from Elsevier.)

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
386 Afferent Synaptic Mechanisms

relatively long (e.g., 5 ms for 100 Hz) compared to the from baseline and evoked SV exocytosis from rat
duration required for fusion of a single vesicle, the IHCs (Glowatzki, E. and Fuchs, P. A., 2002) and
synapse could generate a single spike per cycle if from hair cells of the frog amphibian papilla (Keen,
synaptic exocytosis occurred primarily within a nar- E. C. and Hudspeth, A. J., 2006). In both cases, EPSCs
row time window in response to the onset of each were fast (1–2 ms for the rat, 2–3 ms for the frog) with
depolarizing phase of a periodic stimulus, and was very rapid onset and exponential decay. Sensitivity to
subsequently suppressed until the next depolariza- AMPA receptor antagonists provided a functional
tion. This possibility was tested using capacitance confirmation of previous in vivo and EM studies
measurements as the assay of SV exocytosis (Matsubara, A. et al., 1996), showing that the gluta-
(Rutherford, M. A. and Roberts, W. M., 2006). A matergic receptor subunits GluR2/3 and 4 are
30 ms depolarization to 55 mV was interrupted present.
with a return to 80 mV, thus creating a pair of The amplitude distribution of EPSCs recorded
10 ms depolarizations separated by 10 ms. If the under baseline conditions (i.e., extracellular saline
synapse responds primarily to stimulus onset, then with normal concentration of potassium) showed
the response to the interrupted depolarization is 20-fold variation in the rat and 10-fold variation
expected be twice the response to a maintained in the frog with mean sizes of 190 and 100 pA, respec-
30 ms depolarization, even though the interposed tively (Figure 8). This is remarkable considering that
hyperpolarization reduces the total calcium influx. spontaneous glutamatergic EPSCs recorded elsewhere
The total capacitance change in response to the in the CNS are distributed normally with much smal-
interrupted stimulus was nearly twice as large as ler variation around a quantal size of 20–40 pA.
the response to the sustained depolarization. The Presynaptic depolarization via paired whole-cell
result may reflect the mechanism by which hair cell recordings in the frog, and by perfusion with high-
limits the ANF discharge to a single phase-locked potassium saline in the rat, greatly increased the
spike per stimulus cycle at low frequencies. EPSC rate (Figure 8(a)) while the amplitude distri-
bution remained relatively unchanged (Figures 8(d)
and 8(e)). In both preparations, individual small
3.22.6.2 Encoding of Frequency
events (miniature EPSCs) were observed, but the
In general, all ANFs phase lock to low frequencies, amplitude distributions were skewed to larger events
regardless of their CF. Therefore, the timing of ANF for both baseline and evoked EPSCs. For example, in
spikes can provide information about the frequency of a the rat the peak in the amplitude distribution chan-
low-frequency sound independent of the fiber’s CF ged from about 18 pA in 5.8 mM potassium (baseline,
(Figure 2). For frequencies above 3–4 kHz, the alter- Figure 8(d), left) to about 36 pA in 40 mM potassium
native to temporal coding is place-rate coding (Javel, E. (evoked, Figure 8(e), left), as the mean rate changed
and Mott, J. B., 1988), which does not rely upon the from about 1.5 to 27 events per second. Interestingly,
temporal (fine timing) pattern of spikes to communicate the mean EPSC size was slightly less for evoked
frequency. In the case of the mammalian cochlea, which EPSCs, perhaps due to SV depletion with the sharp
performs exquisite mechanical tuning, frequency can be increase in EPSC frequency. Combined with the
determined from peaks in action potential rates in positive shift in the peak of the amplitude distribu-
ANFs that innervate specific CF regions (labeled tion, this results in a tightening of the EPSC
lines) in the organ of Corti. Therefore, an increase in amplitude distribution (e.g., smaller interquartile
the mean firing rate of an ANF that originates from a range), although the total range remains unchanged
particular place in the organ of Corti is a mechanism for (cf. Figures 8(d) and 8(e), left). In both frog and rat
specifying frequency to the brain. Auditory nuclei in hair cells, the larger EPSCs were sometimes multi-
the brain stem may use different decoding algorithms phasic, but the durations of the larger monophasic
for different frequency ranges in order to interpret events were not appreciably longer than the smaller
firing patterns of the auditory nerve. monophasic events (Figure 8(b), frog; Figures 8(c)–8(e),
rat), which suggests that if the large EPSCs are due to
the fusion of several SVs, then these fusion events are
3.22.6.3 Baseline and Evoked Synaptic
highly synchronized (reviewed in Parsons, T. D. and
Activity
Sterling, P., 2003; Singer, J. H. et al., 2004).
Whole-cell recordings from postsynaptic afferent Two possible mechanisms might explain such multi-
terminals demonstrated EPSCs (Figure 8) resulting quantal events. One possibility is that the opening of

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
Afferent Synaptic Mechanisms 387

(a) –45 mV (b) (c)

300 300

EPSC amplitude (pA)


–40 mV 200 200
200
100 100 pA

0 0
–35 mV 0 0.4 0.8 0 0.4 0.8 1 ms
Rise time (ms) Decay time (ms)

–30 mV (d) 800 AF #4–7


+
5.8 mM K

EPSC amplitude (pA)


20 AF #4–7
n = 698
+ 600
5.8 mM K

Number of events
15 190 ± 149 pA
–25 mV n = 1003
400
10
200
5
–20 mV
0 0
0 200 400 600 800 0.0 0.5 1.0 1.5 2.0
EPSC amplitude (pA) 10–90% rise time (ms)
800
–15 mV (e)

EPSC amplitude (pA)


AF #1
30 AF #1 +
+ 600 40 mM K
40 mM K
n = 340
25 133 ± 126 pA
Number of events

n = 802
–10 mV 20 400
15
10 200
5
20 ms 100 pA 0 0
0 200 400 600 800
0 0.5 1.0 1.5 2.0
EPSC amplitude (pA) 10–90% rise time (ms)

Figure 8 Excitatory postsynaptic currents (EPSCs) in afferent fibers indicate the size and frequency of glutamate release
from hair cells. (a) EPSCs recorded in afferent fibers of the frog amphibian papilla while the presynaptic hair cell was
depolarized to the voltages indicated at the right of each trace. The predominant effect of depolarization was an increase in
EPSC frequency, with little change in EPSC amplitudes. (b) The rise and decay times for baseline EPSCs (frog) are tightly
distributed despite a wide range of amplitudes. (c) EPSCs evoked from rat cochlear afferent fibers in response to inner hair
cell depolarization by 40 mM extracellular potassium, aligned in time, exhibit a large range in the amplitude of monophasic
events. (d, e) Left: EPSC amplitude distributions from rat afferent fibers show the number of events observed in each
amplitude bin along the x-axis. Baseline EPSC amplitudes from four afferents are shown in (d), evoked EPSC amplitudes from
one afferent in (e). The distributions are positively skewed, suggestive of multivesicular release events from individual hair cell
active zones. The range of baseline EPSC amplitudes observed in a normal concentration of extracellular potassium (5.8 mM)
spans the entire range of EPSC amplitudes evoked by depolarization in high potassium (see text). Mean amplitudes were
190 pA (baseline) and 133 pA (evoked). Right: The relationship between EPSC amplitude and 10–90% rise time is very similar
for baseline (d) and evoked (e) events. (a, b) Reproduced from Keen, E. C. and Hudspeth, A. J. 2006. Transfer characteristics
of the hair cell’s afferent synapse. Proc. Natl. Acad. Sci. U. S. A. 103, 5537–5542; Copyright (2006) National Academy of
Sciences, U. S. A. (c–e) Glowatzki, E. and Fuchs, P. A., 2002. Transmitter release at the hair cell ribbon synapse. Nat.
Neurosci. 5, 147–154; Copyright (2002) Nature Publishing Group.

a single presynaptic calcium channel, or a few chan- but the slope of the initial rise can be used instead
nels that happen to be close together, triggers the to measure the EPSP size. In vivo recordings from
synchronized fusion of several nearby vesicles. afferent units in the guinea-pig cochlea demonstrated
Another is that several vesicles undergo nonsynchro- a range of EPSP slopes (Siegel, J. H., 1992). Similar to
nous fusion with each other, but their contents are the EPSCs recorded in vitro and discussed above, the
not released until one of them fuses with the distribution of EPSP slopes was positively skewed,
plasmalemma. suggesting the existence of synchronous multiquantal
When spikes are present in intracellular record- release events. Interestingly, nearly all EPSPs that
ings from ANFs, the peak EPSP amplitude is did not occur during neural refractory periods trig-
obliterated by the action potential that it triggers, gered action potentials.

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
388 Afferent Synaptic Mechanisms

What then is the consequence of multivesicular mean EPSC amplitude did not increase following
release to synaptic function? One possibility is that stimulation (Glowatzki, E. and Fuchs, P. A., 2002;
large EPSCs trigger action potentials with less tem- Keen, E. C. and Hudspeth, A. J., 2006).
poral jitter (Anderson, D. J. et al., 1970; reviewed in
Fuchs, P. A., 2005), but this leaves open the question
3.22.6.4 Tuning Mechanisms in Hair Cells
of whether there is some value in having a large
variance in EPSP amplitude rather than all being Auditory organs of fish, amphibians, reptiles, and
equally large. An intriguing possibility has to do birds all show frequency selectivity resulting from
with the role of efferent innervation on the input the properties of individual hair cells. These mechan-
resistance of afferent terminals. The terminal bouton isms are different from the frequency tuning seen in
of a type I afferent is typically postsynaptic to both an the mammalian cochlea, which is the result of the
IHC and an efferent axon (see Chapter Efferent gradient in stiffness of the basilar membrane (acellu-
System). Spikes in the efferent axon release transmit- lar) and active mechanical feedback from hair bundle
ter that opens postsynaptic channels, thereby motion and OHC motility (cellular). Our focus will
reducing the input resistance of the afferent terminal. be on tuning in the other four classes of vertebrate
The effect of efferent activity may therefore be to animals.
reduce the size of the EPSP produced by a given
EPSC. Since efferent activity increases in loud envir- 3.22.6.4.1 Frequency-selective
onments, the effect may be to render small EPSCs augmentation of membrane potential
from the hair cell unable to trigger ANF spikes in oscillation
loud environments. In this case, the wide distribution In the turtle auditory papilla, there is no apparent
of EPSC amplitudes associated with multiquantal frequency tuning extrinsic to the hair cell (i.e., trans-
events may be important for efferent control of duction currents are equal over a broad frequency
ANF firing rate. Similarly, efferent innervation range). Instead, each cell is intrinsically tuned to a
directly onto hair cells is proposed to evoke inhibi- different frequency band by a process that has been
tory conductances in hair cells during strong stimuli. called electrical resonance. The resonant frequency
The expected effect would be a decrease in hair cell is determined primarily by the number and kinetics
depolarization and calcium influx. Since efferent of the cell’s calcium-activated potassium channels,
modulation of synaptic activity is expected to be voltage-activated potassium and calcium channels,
largely absent from in vitro preparations, the effects and electrical capacitance (Ricci, A. J. et al., 2000;
of such inhibition on the EPSC amplitude and fre- and reviewed in Fettiplace, R. and Fuchs, P. A.,
quency in response to varying presynaptic stimulus 1999). Each hair cell preferentially amplifies the
levels are unknown. receptor potential in a frequency band, with peak
If large EPSCs are due to the fusion of vesicles amplification near the resonant frequency. The tono-
with each other on the SB before one of them fuses topic axis in this organ is generated by the systematic
with the plasmalemma, this could be a mechanism to variation of the hair cells’ preferred frequencies along
integrate sound amplitude over many cycles while the length of the sensory epithelium. Low-frequency
preserving phase locking of glutamate release to the cells have low channel density and slower potassium
sound waveform. For this to work, the fusion of SVs channel kinetics than high-frequency cells. These
with each other on the SB would have to be triggered variations in channel kinetics are caused by differen-
by the relatively small changes in free calcium that tial expression of potassium channel genes and
are predicted to occur at distances of tens to hun- differential splicing of the mRNAs ( Jones, E. M. C.
dreds of nanometers from the plasmalemma, whereas et al., 1999). The result is a continuum of resonant
the fusion of the resulting multivesicular structure frequencies along the papilla. Similar electrical tuning
with the cell surface membrane would be phase has been studied in the frog sacculus (Armstrong, C. E.
locked to the large, rapid calcium changes that and Roberts, W. M., 1998; 2001) and chick cochlea
occur close to the presynaptic calcium channels. In (Correia, M. J., 1992), but there is no evidence of
favor of this hypothesis, large irregularly shaped electrical tuning in the mammalian cochlea.
structures that appear to be attached to the SB are
seen much more frequently at stimulated synapses 3.22.6.4.2 Synaptic frequency selectivity
than inhibited or resting synapses (Lenzi, D. et al., Experiments employing capacitance measurements
1999; 2002, Figure 4). Opposed to this hypothesis, the in hair cells typically use very strong depolarizations

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
Afferent Synaptic Mechanisms 389

that evoke maximal calcium currents in order to 3.22.7 Auditory Nerve Fiber Firing
measure vesicle pool sizes by depletion. This Patterns Mirror Presynaptic Vesicle
approach has been very fruitful for understanding Pool Dynamics
vesicle pools, but the stimuli used for these experi-
ments are much larger than in vivo receptor Until recently, physiological studies of cochlear out-
potentials, drawing into question the physiological put have focused on in vivo recordings of spike trains
relevance of the evoked exocytosis. A recent study in ANFs in response to sound stimuli (e.g., pure tones
of frog saccular hair cells (Rutherford, M. A. and of varying SPL). Advances in patch-clamp techni-
Roberts, W. M., 2006) used smaller, sinusoidal fluc- ques in vitro, such as capacitance measurements as a
tuations in membrane potential as the voltage-clamp presynaptic assay of SV exocytosis, have allowed
stimulus to mimic the size and frequency of in vivo more direct comparisons between presynaptic activ-
receptor potentials and looked for frequency selec- ity and postsynaptic spike trains.
tivity downstream of the known mechanical and
electrical tuning mechanisms in these cells. Changes
in membrane capacitance were measured in response 3.22.7.1 Adaptation of the Auditory Nerve
to small sinusoidal voltage changes (5 mV centered Response
at 60 mV) delivered at 5, 50, and 200 Hz to test the
hypothesis that exocytosis would be largest at 50 Hz, Adaptation of ANF responses is thought to be neces-
the frequency where the sacculus is most sensitive to sary for the peripheral processing of speech
airborne and groundborne vibrations. Consistent (Delgutte, B., 1980). Two types of fast adaptation
with this hypothesis, exocytosis was twice as large have been described to explain the reduction in
ANF discharge rate following the onset of a sustained
at 50 Hz than at 5 or 200 Hz, even though the total
acoustic stimulation (Figure 6(e)). These have been
calcium influx was the same at the three frequencies.
called rapid adaptation and short-term adaptation,
This preference for 50 Hz disappeared when stronger
with durations of approximately 10 and 20–90 ms,
sinusoidal stimuli (10 mV centered at 55 mV)
respectively. Although the precise time courses can
were applied, which is reminiscent of other tuning
vary by species, SR of the fiber, and sound intensity;
mechanisms that are saturated by large stimuli (e.g.,
both processes are thought to arise from mechanisms
de Boer, E. and Nuttall, A. L., 2000). It is important to
downstream of IHC receptor potential in vivo, which
note that the 5 mV sinusoidal stimulus, while small
is not reduced in amplitude during a sustained sti-
compared to the stimuli used in most voltage-clamp
mulus (Russell, I. J. and Sellick, P. M., 1978; Holton,
studies, is well above the amplitude of oscillations T. and Weiss, T. F., 1983). Fast adaptation is thought
that occur at sensory threshold (Dallos, P., 1985). The to involve presynaptic depression (as described
exocytosis in response to the small stimulus at 50 Hz below), but not via inactivation of the presynaptic
corresponded to the net fusion of 1.7 SV per AZ per calcium current which shows no significant decline
stimulus cycle. This amounts to twice the number of over this time course under voltage clamp (Parsons,
vesicles in the docked pool fusing during the 1 s T. D. et al., 1994; Moser, T and Beutner, D., 2000;
stimulus, and suggests that new vesicles can dock Schnee, M. E. and Ricci, A. J., 2003; Eisen, M. D. et al.,
and be primed for release during the stimulus. This 2004).
synaptic tuning, operating in series with mechanical Fast adaptation of ANF firing and its subsequent
and electrical tuning, is expected to contribute to the recovery seem to be predominantly due to depletion
overall tuning of the sacculus, which has been and refilling of the RRP. Recovery from adaptation
measured by action potential firing rates and sensi- takes hundreds of milliseconds (Smith, R. L., 1977;
tivity in afferent axons of the eighth cranial nerve Harris, D. M. and Dallos, P., 1979; Chimento, T. C.
(Koyama, H. et al., 1982; Lewis, E. R. and Narins, P. M., and Schreiner, C. E., 1991), and this time course is
1985; Yu, X. et al., 1991; Christensen-Dalsgaard, J. and thought to reflect refilling of the RRP. In the chick,
Narins, P. M., 1993). Similar synaptic tuning could using a combination of in vivo single unit recordings
contribute to frequency selectivity at other ribbon of ANF activity and in vitro patch-clamp recordings
synapses, particularly in auditory and vestibular of SV fusion, it was found that sound-evoked ANF
organs where synaptic transmission results in activity adapted and recovered with similar time
phased-locked action potentials in the postsynaptic courses as RRP depletion and recovery (Figure 6;
neuron. Spassova, M. A. et al., 2004). After RRP exhaustion,

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
390 Afferent Synaptic Mechanisms

its recovery was 95% complete after a 200 ms rest branch of the eighth nerve of the goldfish and
period, similar to the time course of recovery from observed a 20 ms delay in firing upon a decrement
adaptation. In comparison, analogous recovery in in the intensity of tone. They attributed this delay to
retinal bipolar neurons takes about 100 times longer the time required for SVs to refill the depleted low-
(20 s; von Gersdorff, H. and Matthews, G., 1997). In threshold release sites between the SB and the cal-
the mouse IHC, RRP refilling is biphasic, possibly cium channels at the plasmalemma. Furthermore,
reflecting the refilling of two functionally distinct quantal analysis of the size of EPSPs showed a nega-
types of AZs (Moser, T. and Beutner, D., 2000). tive correlation between the amplitude of two
Differences among AZs in a single IHC could pro- consecutive events (i.e., large EPSPs tend to be fol-
vide the basis for the variable properties of ANFs lowed by smaller ones and vice versa). These data
(Merchan-Perez, A. and Liberman, M. C., 1996). support the idea that adaptive rundown of EPSPs is
Indeed high- and low-SR fibers exhibit different the result of SV depletion (Furukawa, T. and
times needed for recovery from prior stimuli Matsuura, S., 1978; Furukawa, T. et al., 1978).
(Relkin, E. M. and Doucet, J. R., 1991) that are a Additional evidence for the involvement of the SB
close match to the biphasic refilling of the RRP in maintenance of the functional RRP comes from
(Figure 6(d)). the Bassoon mutant mice, in which the first phase of
exocytosis and the calcium current are both reduced
(Khimich, D. et al., 2005). The large reduction of SB-
3.22.7.2 The Role of the Synaptic Body in
associated docked SVs at AZs in the mutant provides
Auditory Nerve Fiber Spike Timing
additional support for the contribution of these
Computational attempts at predicting psychophysi- docked vesicles to the functional RRP measured in
cal performance on auditory tasks based on the capacitance experiments. The reduced calcium cur-
activity of ANFs have generally concluded that the rent supports previous evidence that calcium channel
probabilistic nature of action potentials (the variabil- expression is coregulated with and restricted spa-
ity of spike timing during repeated presentations of tially to SV release sites adjacent to SBs (Martinez-
the same stimulus) limits the precision of acoustic Dunst, C. et al., 1997; Zenisek, D. et al., 2003).
information sent to the brain (Mountain, D. C. and Intriguingly, the sustained component of the capaci-
Hubbard, A. E., 1996). Upon repeated presentation of tance rise was relatively intact in the mutant mice
a pure tone, an ANF will fire a first spike with compared to the 50% reduction in ANF discharge
variable latency. The variance of first spike latency rate after short-term adaptation to tonal stimuli
in mice with a targeted deletion of the gene for the (Buran, B. N. et al., 2006). One interpretation is that
cytomatrix AZ protein Bassoon was increased by a the sustained component in the mutant mice is rela-
factor of 12 compared to controls (Buran, B. N. et al., tively asynchronous or not properly localized to
2006). Bassoon is thought to anchor SBs at AZs by postsynaptic fibers. In contrast, the sustained exocy-
interacting with Ribeye (Khimich, D. et al., 2005; tom tosis in wild-type mice is likely to reflect synaptic
Dieck, S. et al., 2005); ultrastructural inspection release as indicated by the massive sustained release
demonstrated a 90% reduction in the number of measure by EPSCs in high potassium and higher
SBs anchored at AZs. This work suggests that the levels of sustained ANF discharge rate in vivo. The
presence of SBs at release sites helps to reduce the sound-evoked onset firing rate (before adaptation)
temporal jitter of ANF discharge, and hence improve and steady-state firing rate (after fast adaptation)
phase locking. were decreased and threshold was increased, in
In the absence of an applied stimulus, the tem- accordance with the reduction of the functional
poral firing pattern of ANFs in vivo is often RRP measured presynaptically. In spite of this, the
represented as a poison process, but on a short time- shapes of rate-level functions were normal in the
scale it is more accurately described as a renewal mutants. Indeed a recent modeling study showed
process that is a function of neurotransmitter con- that the effect of increasing the maximum number
centration in the cleft (presynaptic effect) and neural of transmitter quanta held in the immediate store
refractoriness (postsynaptic effect), the combination (analogous to the concept of increasing the size of
of which causes the firing probability to be depressed the RRP) is to scale the release rate linearly across
for more than 20 ms following a spike (Gaumond, R. the entire dynamic range (Sumner, C. J. et al., 2002).
P. et al., 1983). Furukawa and co-workers recorded Work with the Bassoon mutant mouse has shown
synaptic potentials from afferent units in the saccular that the localization of synaptic ribbons to hair cell

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
Afferent Synaptic Mechanisms 391

AZs is required for transmitter release to be precisely Armstrong, C. E. and Roberts, W. M. 2001. Rapidly inactivating
and non-inactivating calcium-activated potassium currents
synchronized to the onset of sound at each AZ, which in frog saccular hair cells. J. Physiol. 536, 49–65.
is required for synchronous auditory signaling car- Art, J. J. and Fettiplace, R. 1987. Variation of membrane
ried to the brain stem (Khimich, D. et al., 2005; Buran, properties in hair cells isolated from the turtle cochlea. J.
Physiol. 385, 207–242.
B. N. et al., 2006). If simultaneously activated fibers Art, J. J., Wu, Y. C., and Fettiplace, R. 1995. The calcium-
provide input to a single neuron in the cochlear activated potassium channels of turtle hair cells. J. Gen.
nucleus, then converging synaptic inputs arriving Physiol. 105, 49–72.
Balkema, G. W. 1991. A synaptic antigen (B16) is localized in
closely in time could be the physiological basis for a retinal synaptic ribbons. J. Comp. Neurol. 312, 573–583.
coincidence detector to differentiate between base- Bodian, D. 1980. Presynaptic bodies of auditory hair cells in old
line activity and stimulus-evoked (synchronous) hair world monkeys. Anat. Rec. 197, 379–386.
Bracewell, R. N. 1999. The Fourier Transform and Its
cell afferent synaptic activity. It seems probable that Applications. McGraw-Hill Co.
the precise configurations of SBs and their associated Brandt, A., Khimich, D., and Moser, T. 2005. Few Cav1.3
cytomatrices at AZs are important for establishing channels regulate the exocytosis of a synaptic vesicle at the
hair cell ribbon synapse. J. Neurosci. 25, 11577–11585.
both the sensitivity and the phase relationship of Brandt, A., Striessng, J., and Moser, T. 2003. CaV1.3 channels
ANF spikes to the sound. are essential for development and presynaptic activity of
cochlear inner hair cells. J. Neurosci. 23, 10832–10840.
Bunt, A. H. 1971. Enzymatic digestion of synaptic ribbons in
amphibian retinal photoreceptors. Brain Res. 25, 571–577.
Burack, M. A., Silverman, M. A., and Banker, G. 2000. The role
3.22.8 Synapse to Psychophysics of selective transport in neuronal protein sorting. Neuron
26, 465–474.
One of the goals of auditory research is to provide an Buran, B. N., Moser, T., Gundelfinger, E. D., and Liberman, M.
C. 2006. Role of synaptic ribbons in auditory nerve
explanation of the behavior of the auditory system responses in mice lacking the synaptic scaffolding protein
based upon the functions of its cellular components Bassoon. Assoc. Res. Otolaryngol (ARO) Poster, 2006,
and the circuits they make. For example, numerous Baltimore.
Cheatham, M. A. and Dallos, P. 1993. Longitudinal comparisons
experiments have examined the ability of the audi- of IHC ac and dc receptor potentials recorded from the
tory system to detect weak sounds. These guinea pig cochlea. Hear Res. 68, 107–114.
experiments have been carried at the level of ANFs, Chen, X., Tomchick, D. R., Kovrigin, E., Arac, D., Machius, M.,
Sudhof, T. C., and Rizo, J. 2002. Three-dimensional
cortical neurons, and behavior, in a variety of mam- structure of the complexin/SNARE complex. Neuron
mals, including humans. Recently it has been 33, 397–409.
proposed that the inverse relationship between the Chimento, T. C. and Schreiner, C. E. 1991. Adaptation and
recovery from adaptation in single fiber responses of the cat
duration and the SPL needed to attain a threshold auditory nerve. J. Acoust. Soc. Am. 90, 263–273.
criterion for detection seen at all levels of auditory Chinnadurai, G. 2002. CtBP family proteins: more than
processing can be explained by the properties of hair transcriptional corepressors. Bioessays 25, 9–12.
Christensen-Dalsgaard, J. and Narins, P. M. 1993. Sound and
cell afferent synapses. One interpretation of these vibration sensitivity of VIIIth nerve fibers in the frogs
results is that the statistics of individual events, Leptodactylus albilabris and Rana pipiens pipiens. J. Comp.
which may correspond to the exocytosis of individual Physiol. A 172, 653–662.
Correia, M. J. 1992. Filtering properties of hair cells. Ann. N. Y.
SVs at the afferent synapse, underlies the trade-off Acad. Sci. 656, 49–57.
between strength and duration for barely detectable Dallos, P. 1985. Response characteristics of mammalian
sounds (Heil, P. and Neubauer, H., 2003) that is cochlear hair cells. J. Neurosci. 5, 1591–1608.
Dambly-Chaudiere, C., Sapede, D., Soubiran, F., Decorde, K.,
measured in psychophysical experiments of sound Gompel, N., and Ghysen, A. 2003. The lateral line of
perception. zebrafish: a model system for the analysis of morphogenesis
and neural development in vertebrates. Biol. Cell
95, 579–587.
Dargan, S. L. and Parker, I. 2003. Buffer kinetics shape the
spatiotemporal patterns of IP3 evoked Ca2þ signals. J.
References Physiol. 553, 775–788.
de Boer, E. and Nuttall, A. L. 2000. The mechanical waveform of
Achanzar, W. E. and Ward, S. 1997. A nematode gene required the basilar membrane. III. Intensity effects. J. Acoust. Soc.
for sperm vesicle fusion. J. Cell Sci. 110, 1073–1081. Am. 107, 1497–1507.
Anderson, D. J., Rose, J. E., Hind, J. E., and Brugge, J. F. 1970. Delcomyn, F. 1998. Foundations of Neurobiology. W. H.
Temporal position of discharges in single auditory nerve Freeman.
fibers within the cycle of a sine-wave stimulus: frequency Delgutte, B. 1980. Representation of speech-like sounds in the
and intensity effects. J. Acoust. Soc. Am. 49, 1131–1139. discharge patterns of auditory-verve fibers. J. Acoust. Soc.
Armstrong, C. E. and Roberts, W. M. 1998. Electrical properties Am. 68, 843–857.
of frog saccular hair cells: distortion by enzymatic tom Dieck, S., Altrock, W. D., Kessels, M. M., Qualmann, B.,
dissociation. J. Neurosci. 18, 2962–2973. Regus, H., Brauner, D., Fejtova, A., Bracko, O.,

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
392 Afferent Synaptic Mechanisms

Gundelfinger, E. D., and Brandstatter, J. H. 2005. Gray, E. G. and Pease, H. L. 1971. On understanding the origins
Molecular dissection of the photoreceptor ribbon synapse: of the retinal receptor synapses. Brain Res. 35, 1–15.
physical interaction of Bassoon and Ribeye is essential for Griesinger, C. B., Richards, C. D., and Ashmore, J. F. 2005. Fast
the assembly of the ribbon complex. J. Cell Biol. vesicle replenishment allows indefatigable signaling at the
168, 825–836. first auditory synapse. Nature 435, 212–215.
Dobrunz, L. E. and Stevens, C. F. 1997. Heterogeneity of Hackney, C. M., Mahendrasingam, S., Penn, A., and
release probability, facilitation, and depletion at central Fettiplace, R. 2005. The concentrations of calcium buffering
synapses. Neuron 18, 995–1008. proteins in mammalian cochlear hair cells. J. Neurosci.
Edmonds, B. W., Gregory, F. D., and Schweizer, F. E. 2004. 25, 7867–7886.
Evidence that fast exocytosis can be predominantly Hakuba, N., Koga, K., Gyo, K., Usami, S., and Tanaka, K. 2000.
mediated by vesicles not docked at active zones in frog Exacerbation of noise-induced hearing loss in mice lacking
saccular hair cells. J. Physiol. 560, 439–450. the glutamate transporter GLAST. J. Neurosci.
Edmonds, B., Reyes, R., Schwaller, B., and Roberts, W. M. 20, 8750–8753.
2000. Calretinin modifies presynaptic calcium signaling in Hama, K. and Saito, K. 1977. Fine structure of the afferent
frog saccular hair cells. Nat. Neurosci. 3, 786–790. synapse of the hair cells in the saccular macula of the
Eisen, M. D., Spassova, M., and Parsons, T. D. 2004. Large goldfish, with special reference to the anastomosing tubules.
releasable pool of synaptic vesicles in chick cochlear hair J. Neurocytol. 6, 361–373.
cells. J. Neurophysiol. 91, 2422–2428. Harris, D. M. and Dallos, P. 1979. Forward masking of auditory
Fettiplace, R. and Fuchs, P. A. 1999. Mechanisms of hair cell nerve fiber responses. J. Neurophysiol. 42, 1083–1107.
tuning. Annu. Rev. Physiol. 61, 809–834. Hashimoto, S., Kimura, R. S., and Takasaka, T. 1990.
Fuchs, P. A. 2005. Time and intensity coding at the hair cell’s Computer-aided 3-dimensional reconstruction of the inner
ribbon synapse. J. Physiol. 566, 7–12. hair-cells and their nerve-endings in the guinea-pig. Acta
Fuchs, P. A., Glowatzki, E., and Moser, T. 2003. The afferent Otolaryngol. 109, 228–234.
synapse of cochlear hair cells. Curr. Opin. Neurobiol. Heidelberger, R., Sterling, P., and Matthews, G. 2002. Roles of
13, 452–458. ATP in depletion and replenishment of the releasable pool of
Furness, D. N. and Lawton, D. M. 2003. Comparative synaptic vesicles. J. Neurophysiol. 88, 98–106.
distribution of glutamate transporters and receptors in Heil, P. and Neubauer, H. 2001. Temporal integration of sound
relation to afferent innervation density in the mammalian pressure determines thresholds of auditory nerve fibers. J.
cochlea. J. Neurosci. 23, 11296–11304. Neurosci. 21, 7404–7415.
Furness, D. N. and Lehre, K. P. 1997. Immunocytochemical Heil, P. and Neubauer, H. 2003. A unifying basis of auditory
localization of a high-affinity glutamate-aspartate thresholds based on temporal summation. Proc. Natl. Acad.
transporter, GLAST, in the rat and guinea-pig cochlea. Eur. J. Sci. U. S. A. 100, 6151–6156.
Neurosci. 9, 1961–1969. Heller, S., Bell, A. M., Charlotte, S. D., Choe, Y., and
Furukawa, T. and Matsuura, S. 1978. Adaptive rundown of Hudspeth, A. J. 2002. Parvalbumin 3 is an abundant Ca2þ
excitatory post-synaptic potentials at synapses between buffer in hair cells. JARO 3, 488–498.
hair cells and eighth nerve fibers in the goldfish. J. Physiol. Hilfiker, S., Pieribone, V. A., Czernik, A. J., Kao, H. T.,
276, 193–209. Augustine, G. J., and Greengard, P. 1999. Synapsins as
Furukawa, T., Hayashida, Y., and Matsuura, S. 1978. Quantal regulators of neurotransmitter release. Phil. Trans. R Soc.
analysis of the size of excitatory post-synaptic potentials at Lond. B 354, 269–279.
synapses between hair cells and afferent nerve fibers in the Hille, B. 2001. Ionic Channels of Excitable Membranes. Sinauer
goldfish. J. Physiol. 276, 211–226. Associates.
Furukawa, T., Kuno, M., and Matsuura, S. 1982. Quantal Hirokawa, N., Sobue, K., Kanda, K., Harada, A., and Yorifuji, H.
analysis of a decremental response at hair cell-afferent fibre 1989. The cytoskeletal architecture of the presynaptic
synapses in the goldfish sacculus. J. Physiol. 322, 181–195. terminal and molecular structure of synapsin 1. J. Cell Biol.
Gaumond, R. P., Kim, D. O., and Molnar, C. E. 1983. Response of 108, 111–126.
cochlear nerve fibers to brief acoustic stimuli: role of Holt, M., Cooke, A., Neef, A., and Lagnado, L. 2004. High
discharge-history effects. J. Acoust. Soc. Am. 74, 1392–1398. mobility of vesicles supports continuous exocytosis at a
von Gersdorff, H. and Matthews, G. 1997. Depletion and ribbon synapse. Curr. Biol. 14, 173–183.
replenishment of vesicle pools at a ribbon-type synaptic Holt, J. C., Xue, J. T., Brichta, A. M., and Goldberg, J. M.
terminal. J. Neurosci. 17, 1919–1927. 2006. Transmission between type II hair cells and bouton
von Gersdorff, H., Sakaba, T., Berglund, K., and Tachibana, M. afferents in the turtle posterior crista. J. Neurophysiol.
1998. Submillisecond kinetics of glutamate release from a 95, 428–452.
sensory synapse. Neuron 21, 1177–1188. Holton, T. and Weiss, T. F. 1983. Receptor potentials of lizard
Giannessi, F. and Ruffoli, R. 1996. Fine structure of the afferent cochlear hair cells with free-standing stereocilia in response
synapses in the paratympanic organ of the chicken, with to tones. J. Physiol. 345, 205–240.
special reference to the synaptic bodies. Ann. Anat. Hossain, W. A., Antic, S. D., Yang, Y., Rasband, M. N., and
178, 127–131. Morest, D. K. 2005. Where is the spike generator of the
Gitler, D., Takagishi, Y., Feng, J., Ren, Y., Rodriguiz, R. M., cochlear nerve? J. Neurosci. 25, 6857–6868.
Wetsel, W. C., Greengard, P., and Augustine, G. J. 2004. Indresano, A. A., Frank, J. E., Middleton, P., and Jaramillo, F.
Different presynaptic roles of synapsins at excitatory and 2004. Mechanical noise enhances signal transmission in the
inhibitory synapses. J. Neurosci. 24, 11368–11380. bullfrog sacculus. JARO 4, 363–370.
Gleisner, L., Flock, A., and Wersall, J. 1973. The ultrastructure of Issa, N. P. and Hudspeth, A. J. 1994. Clustering of Ca2þ
the afferent synapse on hair cells in the frog labyrinth. Acta channels and Ca2þ-activated Kþ channels at fluorescently
Otolaryngol. 76, 199–207. labeled presynaptic active zones of hair cells. Proc. Natl.
Glowatzki, E. and Fuchs, P. A. 2002. Transmitter release at the Acad. Sci. U. S. A. 91, 7578–7582.
hair cell ribbon synapse. Nat. Neurosci. 5, 147–154. Jacobs, R. A. and Hudspeth, A. J. 1990. Ultrastructural
Goldberg, J. M. 1996. Theoretical analysis of intercellular correlates of mechanoelectrical transduction in hair-cells of
communication between the vestibular type I hair cell and its the bullfrogs internal ear. Cold Spring Harb. Symp. Quant.
calyx ending. J. Neurophysiol. 76, 1942–1957. Biol. 55, 547–561.

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
Afferent Synaptic Mechanisms 393

Javel, E. and Mott, J. B. 1988. Physiological and Lewis, E. R., Baird, R. A., Leverenz, E. L., and Koyama, H. 1982.
psychophysical correlates of temporal processes in hearing. Inner ear: dye injection reveals peripheral origins of specific
Hear Res. 34, 275–294. sensitivities. Science 215, 1641–1643.
Jeffress, L. A. 1968. Mathematical and electrical models of Liberman, M. C. 1980. Morphological differences among radial
auditory detection. J. Acoust. Soc. Am. 44, 187–203. afferent fibers in the cat cochlea: an electron-microscopic
Johnson, D. H. 1980. The relationship between spike rate and study of serial sections. Hear Res. 3, 45–63.
synchrony in responses of auditory-nerve fibers to single Liberman, M. C. 1982. Single-neuron labeling of the cat auditory
tones. J. Acoust. Soc. Am. 68, 1115–1122. nerve. Science 216, 1239–1241.
Johnson, D. H. and Kiang, N. Y. S. 1976. Analysis of discharges Liberman, M. C. and Kiang, N. Y. S. 1978. Acoustic trauma in
recorded simultaneously from pairs of auditory nerve fibers. cats. Cochlear pathology and auditory-nerve activity. Acta
Biophys. J. 16, 719–734. Otolaryngol. Suppl. 358, 1–63.
Johnson, S. L., Marcotti, W., and Kros, C. J. 2005. Increase in Liberman, M. C., Dodds, L. W., and Pierce, S. 1990. Afferent
efficiency and reduction in Ca2þ dependence of exocytosis and efferent innervation of the cat cochlea: quantitative
during development of mouse inner hair cells. J. Physiol. analysis with light and electron microscopy. J. Comp.
563, 177–191. Neurol. 301, 443–460.
Jones, E. M. C., Gray-Keller, M., and Fettiplace, R. 1999. The Lindau, M. and Neher, E. 1988. Patch-clamp techniques for
role of Ca2þ-activated Kþ channel spliced variants in the time-resolved capacitance measurements in single cells.
tonotopic organization of the turtle cochlea. J. Physiol. Pflugers Arch. 411, 137–146.
518, 653–665. Lioudyno, M., Hiel, H., Kong, J. H., Katz, E., Waldman, E.,
Kandel, E. R., Schwartz, J. H., and Jessell, T. M. 2000. Parameshwaran-Iyer, S., Glowatzki, E., and Fuchs, P. A.
Principles of Neural Science. McGraw-Hill. 2004. A ‘‘synaptoplasmic cistern’’ mediates rapid inhibition
Keen, E. C. and Hudspeth, A. J. 2006. Transfer characteristics of cochlear hair cells. J. Neurosci. 24, 11160–11164.
of the hair cell’s afferent synapse. Proc. Natl. Acad. Sci. U. S. Manley, G. A., Köppl, C., and Yates, G. K. 1997. Activity of
A. 103, 5537–5542. primary auditory neurons in the cochlear ganglion of the emu
Khimich, D., Nouvian, R., Pujol, R., tom Dieck, S., Egner, A., Dromaius novaehollandiae: spontaneous discharge,
Gundelfinger, E. D., and Moser, T. 2005. Hair cell synaptic frequency tuning, and phase locking. J. Acoust. Soc. Am.
ribbons are essential for synchronous auditory signalling. 101, 1560–1573.
Nature 434, 889–894. Martinez-Dunst, C., Michaels, R. L., and Fuchs, P. A. 1997.
Kiang, N. Y. S., Rho, J. M., Northrop, C. C., Liberman, M. C., Release sites and calcium channels in hair cells of the
and Ryugo, D. K. 1982. Hair-cell innervation by spiral chick’s cochlea. J. Neurosci. 17, 9133–9144.
ganglion cells in adult cats. Science 217, 175–177. Matsubara, A., Laake, J. H., Davanger, S., Usami, S., and
Kiang, N. Y. S., Watanabe, T., Thomas, E. C., and Clark, L. F. Ottersen, O. P. 1996. Organization of AMPA receptor
1965. Discharge Pattern of Single Fibers in the Cat’s subunits at a glutamate synapse: a quantitative immunogold
Auditory Nerve. MIT Press. analysis of hair cell synapses in the rat organ of Corti. J.
King, T. S. and Dougherty, W. J. 1982. Effect of denervation on Neurosci. 16, 4457–4467.
‘‘synaptic’’ ribbon population in the rat pineal gland. J. McMahon, H. T., Missler, M., Li, C., and Sudhof, T. C. 1995.
Neurocytol. 11, 19–28. Complexins: cytosolic proteins that regulate SNAP receptor
Knudsen, E. I. and Konishi, M. 1978. A neural map of auditory function. Cell 83, 111–119.
space in the owl. Science 200, 795–797. Merchan-Perez, A. and Liberman, M. C. 1996. Ultrastructural
Koschak, A., Reimer, D., Huber, I., Grabner, M., Glossmann, H., differences among afferent synapses on cochlear hair cells:
Engel, J., and Striessnig, J. 2001. 1D (Cav1.3) subunits can correlations with spontaneous discharge rate. J. Comp.
form L-type Ca2þ channels activating at negative voltages. J. Neurol. 371, 208–221.
Biol. Chem. 276, 22100–22106. Miller, M. R. and Beck, J. 1987. Heterotopic synaptic bodies in
Koyama, H., Lewis, E. R., Leverenz, E. L., and Baird, R. A. 1982. the auditory hair cells of adult lizards. Anat. Rec. 218, 338–344.
Acute seismic sensitivity in the bullfrog ear. Brain Res. Morgans, C. W. 2001. Localization of the 1F calcium channel
250, 168–172. subunit in the rat retina. Invest. Ophthalmol. Vis. Sci.
von Kriegstein, K., Schmitz, F., Link, E., and Sudhof, T. C. 42, 2414–2418.
1999. Distribution of synaptic vesicle proteins in the Morgans, C. W., Brandstatter, J. H., Kellerman, J., Betz, H., and
mammalian retina identifies obligatory and facultative Wassle, H. 1996. SNARE complex containing syntaxin 3 is
components of ribbon synapses. Eur. J. Neurosci. present in ribbon synapses of the retina. J. Neurosci.
11, 1335–1348. 16, 6713–6721.
Krstic, R. 1976. Ultracytochemistry of the synaptic ribbons in Moser, T. and Beutner, D. 2000. Kinetics of exocytosis and
the rat pineal organ. Cell Tissue Res. 166, 135–143. endocytosis at the cochlear inner hair cell afferent
Lenzi, D. and von Gersdorff, H. 2001. Structure suggests synapse of the mouse. Proc. Natl. Acad. Sci. U. S. A.
function: the case for synaptic ribbons as exocytotic 97, 883–888.
nanomachines. Bioessays 23, 831–840. Moser, T., Brandt, A., and Lysakowski, A. 2006. Hair cell ribbon
Lenzi, D., Crum, J., Ellisman, M. H., and Roberts, W. M. 2002. synapses. Cell Tissue Res. 326, 347–359.
Depolarization redistributes synaptic membrane and creates Moser, T., Neef, A., and Khimich, D. 2006. Mechanisms
a gradient of vesicles on the synaptic body at a ribbon underlying the temporal precision of sound coding at the
synapse. Neuron 36, 649–659. inner hair cell ribbon synapse. J. Physiol. 576, 55–62.
Lenzi, D., Runyeon, J. W., Crum, J., Ellisman, M. H., and Mountain, D. C. and Hubbard, A. E. 1996. Computational
Roberts, W. M. 1999. Synaptic vesicle populations in Analysis of Hair Cell and Auditory Nerve Processes.
saccular hair cells reconstructed by electron tomography. J. In: Auditory Computation (eds. H. L. Hawkins, T. A. McMullen,
Neurosci. 19, 119–132. A. N. Popper, and R. R. Fay), p. 121. Springer.
Lewis, E. R. and Narins, P. M. 1985. Do frogs communicate with Muresan, V., Lyass, A., and Schnapp, B. J. 1999. The kinesin
seismic signals? Science 227, 187–189. motor KIF3A is a component of the presynaptic ribbon in
Lewis, R. S. and Hudspeth, A. J. 1983. Voltage- and ion- vertebrate photoreceptors. J. Neurosci. 19, 1027–1037.
dependent conductances in solitary vertebrate hair cells. Nakagawa, T., Oghalai, J. S., Saggau, P., Rabbitt, R. D., and
Nature 304, 538–541. Brownell, W. E. 2006. Photometric recording of

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
394 Afferent Synaptic Mechanisms

transmembrane potential in outer hair cells. J. Neural. Eng. Ricci, A. J., Gray-Keller, M., and Fettiplace, R. 2000. Tonotopic
3, 79–86. variations of calcium signalling in turtle auditory hair cells. J.
Neher, E. and Augustine, G. J. 1992. Calcium gradients Physiol. 524, 423–436.
and uffers in bovine chromaffin cells. J. Physiol. Rizzoli, S. O. and Betz, W. J. 2005. Synaptic vesicle pools. Nat.
450, 273–301. Rev. Neurosci. 6, 57–69.
Nguyen, T. H. and Balkema, G. W. 1999. Antigenic epitopes of Roberts, W. M. 1993. Spatial calcium buffering in saccular hair
the photoreceptor synaptic ribbon. J. Comp. Neurol. cells. Nature 363, 74–76.
413, 209–218. Roberts, W. M. 1994. Localization of calcium signals by a
Nibu, Y., Zhang, H., Bajor, E., Barolo, S., Small, S., and mobile calcium buffer in frog saccular hair cells. J. Neurosci.
Levine, M. 1998. dCtBP mediates transcriptional repression 14, 3246–3262.
by Knirps, Krueppel and Snail in the Drosophila embryo. Roberts, W. M., Jacobs, R. A., and Hudspeth, A. J. 1990.
EMBO J. 17, 7009–7020. Colocalization of ion channels involved in frequency
Nicholls, J. G., Martin, A. R., Wallace, B. G., and Fuchs, P. A. selectivity and synaptic transmission at presynaptic active
2001. From Neuron to Brain. Sinauer Associates. zone of hair cells. J. Neurosci. 10, 3664–3684.
Nouvian, R., Beutner, D., Parsons, T. D., and Moser, T. 2006. Robertson, D. 1984. Horseradish peroxidase injection of
Structure and function of the hair cell ribbon synapse. J. physiologically characterized afferent and efferent neurons
Memb. Biol. 209, 1–13. in the guinea pig spiral ganglion. Hear Res. 15, 113–121.
Osborne, M. P. and Thornhill, R. A. 1972. The effect of Robertson, D. and Paki, B. 2002. A role for purinergic receptors
monoamine depleting drugs upon the synaptic bars in the at the inner hair cell-afferent synapse? Audiol. Neurootol.
inner ear of the bullfrog. Z. Zellforsch. Mikrosk. Anat. 7, 62–67.
127, 347–355. Rodriguez-Contreras, A. and Yamoah, E. N. 2001. Direct
Pabst, S., Hazzard, J. W., Antonin, W., Sudhof, T. C., Jahn, R., measurement of single-channel Ca2þ currents in bullfrog hair
Rizo, J., and Fasshauer, D. 2000. Selective interaction of cells reveals two distinct channel subtypes. J. Physiol.
complexin with the neuronal SNARE complex. J. Biol. Chem. 534, 669–689.
275, 19808–19818. Rodriguez-Contreras, A. and Yamoah, E. N. 2003. Effects of
Palmer, A. R. and Russell, I. J. 1986. Phase-locking in the permeant ion concentrations on the gating of L-type Ca2þ
cochlear nerve of the guinea-pig and its relation to the channels in hair cells. Biophys. J. 84, 3457–3469.
receptor potential of inner hair-cells. Hear Res. 24, 1–15. Rosahl, T., Spillane, D., Missler, M., Herz, J., Selig, D.,
Parsons, T. D. and Sterling, P. 2003. Synaptic ribbon: conveyor Wolff, J. R., Hammer, R. E., Malenka, R. C., and
belt or safety belt? Neuron 37, 379–82. Sudhof, T. C. 1995. Essential functions of synapsins I and II
Parsons, T. D., Lenzi, D., Almers, W., and Roberts, W. M. 1994. in synaptic vesicle regulation. Nature 375, 488–493.
Calcium-triggered exocytosis and endocytosis in an isolated Rose, J. E., Brugge, J. F., Anderson, D. J., and Hind, J. E. 1967.
presynaptic cell: capacitance measurements in saccular hair Phase-locked response to low-frequency tones in single
cells. Neuron 13, 875–883. auditory nerve fibers of the squirrel monkey. J. Neurophysiol.
Plack, C. J., Oxenham, A. J., and Drga, V. 2006. Masking by 30, 769–793.
inaudible sounds and the linearity of temporal summation. J. Rose, J. E., Hind, J. E., Anderson, D. J., and Brugge, J. F. 1971.
Neurosci. 26, 8767–8773. Some effects of stimulus intensity on response of auditory-
Poortinga, G., Watanabe, M., and Parkhurst, S. M. 1998. nerve fibers in the squirrel monkey. J. Neurophysiol.
Drosophila CtBP: a hairy-interacting protein required for 34, 685–699.
embryonic segmentation and hairy-mediated transcriptional Rosenmund, C., Rettig, J., and Brose, N. 2003. Molecular
repression. EMBO J. 17, 2067–2078. mechanisms of active zone function. Curr. Opin. Neurobiol.
Pyle, J. L., Kavalali, E. T., Piedras-Renteria, E. S., and 13, 509–519.
Tsien, R. W. 2000. Rapid reuse of readily releasable pool Ruel, J., Bobbin, R. P., Vidal, D., Pujol, R., and Puel, J. L. 2000.
vesicles at hippocampal synapses. Neuron 28, 221–231. The selective AMPA receptor antagonist GYKI 53784 blocks
Raphael, Y. and Altschuler, R. A. 2003. Structure and action potential generation and excitotoxicity in the guinea
innervation of the cochlea. Brain Res. Bull. 60, 397–422. pig cochlea. Neuropharmacology 39, 1959–1973.
Raymond, J., Puel, J. L., and Devau, G. 2003. Glutamate Russell, I. J. and Sellick, P. M. 1978. Intracellular studies of hair
transporters in the guinea-pig cochlea: partial mRNA cells in the mammalian cochlea. J. Physiol. 284, 261–290.
sequences, cellular expression and functional implications. Rutherford, M. A. and Roberts, W. M. 2006. Frequency
Eur. J. Neurosci. 17, 83–92. selectivity of exocytosis in frog saccular hair cells. Proc. Natl.
Rea, R., Li, J., Dharia, A., Levitan, E. S., Sterling, P., and Acad. Sci. U. S. A. 103, 2898–2903.
Kramer, R. H. 2004. Streamlined synaptic vesicle cycle in Safieddine, S. and Wenthold, R. J. 1999. SNARE complex at the
cone photoreceptor terminals. Neuron 41, 755–766. ribbon synapses of cochlear hair cells: analysis of synaptic
Rebillard, G., Ruel, J., Nouvian, R., Saleh, H., Pujol, R., vesicle- and synaptic membrane-associated proteins. Eur. J.
Dehnes, Y., Raymond, J., Puel, J. L., and Devau, G. 2003. Neurosci. 11, 803–812.
Glutamate transporters in the guinea-pig cochlea: partial Schmitz, F. and Drenckhahn, D. 1991. Influence of Ca2þ on
mRNA sequences, cellular expression and functional synaptic morphology of fish cone photoreceptors. Brain Res.
implications. Eur J. Neurosci. 17, 83–92. 546, 341–344.
Reim, K., Mansour, M., Varoqueaux, F., McMahon, H. T., Schmitz, F., Konigstorfer, A., and Sudhof, T. C. 2000. RIBEYE, a
Sudhof, T. C., Brose, N., and Rosenmund, C. 2001. component of synaptic ribbons: a protein’s journey through
Complexins regulate a late step in Ca2þ-dependent evolution provides insight into synaptic ribbon function.
neurotransmitter release. Cell 104, 71–81. Neuron 28, 857–872.
Reim, K., Wegmeyer, H., Brandstatter, J. H., Xue, M., Schnee, M. E. and Ricci, A. J. 2003. Biophysical and
Rosenmund, C., Dresbach, T., Hofmann, K., and Brose, N. pharmacological characterization of voltage-gated calcium
2006. Structurally and functionally unique complexins at currents in turtle auditory hair cells. J. Physiol. 549, 697–717.
retinal ribbon synapses. J. Cell Biol. 169, 669–680. Schnee, M. E., Lawton, D. M., Furness, D. N., Benke, T. A., and
Relkin, E. M. and Doucet, J. R. 1991. Recovery from prior Ricci, A. J. 2005. Auditory hair cell-afferent fiber synapses
stimulation. I. Relationship to spontaneous firing rates of are specialized to operate at their best frequencies. Neuron
primary auditory neurons. Hear Res. 55, 215–222. 47, 243–254.

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395


Author's personal copy
Afferent Synaptic Mechanisms 395

Schneggenburger, R. and Neher, E. 2005. Presynaptic calcium Stevens, C. F. and Williams, J. H. 2000. ‘‘Kiss and run’’
and control of vesicle fusion. Curr. Opin. Neurobiol. exocytosis at hippocampal synapses. Proc. Natl. Acad. Sci.
15, 266–274. U. S. A. 97, 12828–12833.
Sewell, W. F. 1984. The relation between the endocochlear Sudhof, T. C. 2000. The synaptic vesicle cycle revisited. Neuron
potential and spontaneous activity in auditory nerve fibres of 28, 317–320.
the cat. J. Physiol. 347, 685–696. Sudhof, T. C. 2004. The synaptic vesicle cycle. Annu. Rev.
Siegel, J. H. 1992. Spontaneous synaptic potentials from Neurosci. 27, 509–547.
afferent terminals in the guinea pig cochlea. Hear Res. Sumner, C. J., Lopez-Poveda, E. A., O’Mard, L. P., and
59, 85–92. Meddis, R. 2002. A revised model of the inner-hair cell and
Siegel, J. H. and Brownell, W. E. 1981. Synaptic bodies in outer auditory nerve complex. J. Acoust. Soc. Am. 111, 2178–2188.
hair cells of the chinchilla organ of Corti. Brain Res. Tsuji, J. and Liberman, M. C. 1997. Intracellular labeling of
220, 188–193. auditory nerve fibers in the guinea pig: central and peripheral
Siegel, J. H. and Brownell, W. E. 1986. Synaptic and golgi projections. J. Comp. Neurol. 381, 188–202.
membrane recycling in cochlear hair cells. J. Neurocytol. Usukura, J. and Yamada, E. 1987. Ultrastructure of the synaptic
15, 311–328. ribbons in photoreceptor cells of Rana catesbeiana revealed
Singer, J. H., Lassova, L., Vardi, N., and Diamond, J. S. 2004. by freeze-etching and freeze-substitution. Cell Tissue Res.
Coordinated multivesicular release at a mammalian ribbon 247, 483–488.
synapse. Nat. Neurosci. 7, 826–833. Wagner, H.-J. 1978. Cell types and connectivity patterns in
Sjöstrand, F. S. 1958. Ultrastructure of retinal rod synapses of mosaic retinas. Adv. Anat. Embryol. Cell Biol. 55, 1–81.
the guinea pig eye as revealed by three-dimensional Wagner, H.-J. 1997. Presynaptic bodies (‘‘ribbons’’): from
reconstructions from serial sections. J. Ultrastruct. Res. ultrastructural observations to molecular perspectives. Cell
5, 13–17. Tissue Res. 287, 435–446.
Slepecky, N. B., Galsky, M. D., Swartzentruber-Martin, H., and Wan, L., Almers, W., and Chen, W. 2005. Two ribeye genes in
Savage, J. 2000. Study of afferent nerve terminals and fibers in teleosts: the role of Ribeye in ribbon formation and bipolar
the gerbil cochlea: distribution by size. Hear Res. 144, 124–134. cell development. J. Neurosci. 25, 941–994.
Smith, R. L. 1977. Short-term adaptation in single auditory Xu, W. and Lipscombe, D. 2001. Neuronal Cav1.3 1 L-type
nerve fibers: some poststimulatory effects. J. Neurophysiol. channels activate at relatively hyperpolarized membrane
40, 1098–1111. potentials and are incompletely inhibited by
Smith, C. A. and Sjostrand, F. S. 1961. Structure of the nerve dihydropyridines. J. Neurosci. 21, 5944–5951.
endings on the external hair cells of the guinea pig cochlea Yamazaki, H., Nakata, T., Okada, Y., and Hirokawa, N. 1995.
as studied by serial sections. J. Ultrastruct. Res. 5, 523556. KIF3A/B: a heterodimeric kinesin superfamily protein that
Smotherman, M. S. and Narins, P. M. 2000. Hair cells, hearing works as a microtubule plus end-directed motor for
and hopping: a field guide to hair cell physiology in the frog. membrane organelle transport. J. Cell Biol. 130, 1387–1399.
J. Exp. Biol. 203, 2237–2246. Yasunaga, S., Grati, M., Chardenoux, S., Smith, T. N.,
Sneary, M. G. 1988. Auditory receptor of the red-eared turtle: II. Friedman, T. B., Lalwani, A. K., Wilcox, E. R., and Petit, C.
Afferent and efferent synapses and innervation patterns. J. 2000. OTOF encodes multiple long and short isoforms:
Comp. Neurol. 276, 588–606. genetic evidence that the long ones underlie recessive
Sobkowicz, H. M., August, B. K., and Slapnick, S. M. 2004. deafness DFNB9. Am. J. Hum. Genet. 67, 591–600.
Synaptic arrangements between inner hair cells and tunnel Yasunaga, S., Grati, M., Cohen-Salmon, M., El-Amraoui, A.,
fibers in the mouse cochlea. Synapse 52, 299–315. Mustapha, M., Salem, N., El-Zir, E., Loiselet, J., and Petit, C.
Sobkowicz, H. M., Rose, J. E., Scott, G. L., and Levenick, C. V. 1999. A mutation in OTOF, encoding Otoferlin, a FER-1-like
1986. Distribution of synaptic ribbons in the developing protein, causes DFNB9, a nonsyndromic form of deafness.
organ of Corti. J. Neurocytol. 15, 693–714. Nat. Gen. 21, 363–369.
Sobkowicz, H. M., Slapnick, S. M., and August, B. K. 2003. Yu, X., Lewis, E. R., and Feld, D. 1991. Seismic and auditory
Reciprocal synapses between inner hair cell spines and tuning curves from bullfrog saccular and amphibian papillar
afferent dendrites in the organ of corti of the mouse. Synapse axons. J. Comp. Physiol. A 169, 241–248.
50, 53–66. Zenisek, D., Davila, V., Wan, L., and Almers, W. 2003. Imaging
Spassova, M. A., Avissar, M., Furman, A. C., Crumling, M. A., calcium entry sites and ribbon structures in two presynaptic
Saunders, J. C., and Parsons, T. D. 2004. Evidence that cells. J. Neurosci. 23, 2538–2548.
rapid vesicle replenishment of the synaptic ribbon mediates Zhai, R. G. and Bellen, H. J. 2004. The architecture of the active
recovery from short-term adaptation at the hair cell afferent zone in the presynaptic nerve terminal. Physiology 19, 262–270.
synapse. JARO 5, 376390.
Spoendlin, H. 1969. Innervation patterns in the organ of Corti of
the cat. Acta Otolaryngol. 67, 239–254.
Spoendlin, H. 1985. Anatomy of cochlear innervation. Am. J.
Otolaryngol. 6, 453–467. Further Reading
Sterling, P. and Matthews, G. 2005. Structure and function of
ribbon synapses. Trends Neurosci. 28, 20–29. Rizzoli, S. O. and Betz, W. J. 2004. The structural organization
Stern, M. D. 1992. Buffering of calcium in the vicinity of a of the readily releasable pool of synaptic vesicles. Science
channel pore. Cell Calcium 13, 183–192. 303, 2037–2039.

The Senses: A Comprehensive Reference, vol. 3, pp. 365-395

You might also like