You are on page 1of 46

4

Phys i o l ogy of the


Vesti bu l ar Organs

The vestibular organs monitor the motion 4.1 G E N E RAL FEATU RES Of THE
of the head and the forces acting on it. As we VESTI B U LAR ORGANS
have noted in earlier chapters, each semicircular
canal measures the angular motion of the head
Vest i b u lar O rga n s Are I n ertial S e n s o rs
around a single axis, while the three canals, given
their nearly orthogonal orientation, provide the A critical event in transduction is the bending
brain with a three-dimensional reconstruction of hair bundles as a result of the displacement
of this motion. Similarly, the two otolith organs of a gelatinous accessory structure, either
furnish a three-dimensional reconstruction the cupula or otoconial membrane, relative to
of translational motion. In addition, the otolith the apical surface of the neuroepithelium.
organs, because of their sensitivity to linear Displacement takes place on three space scales.
forces, respond to tilts that change the orienta­ Macromechanics refers to the bulk movement of
tion of the head with respect to the earth's gravi­ the accessory structure, together with the associ­
tational field. ated movement of other labyrinthine structures
Given the differences in their design and and fluids. The coupling of hair bundles to the
function, we will consider the canals and otolith accessory structure is referred to as microme­
organs separately. Before doing so, however, we chanics, while the linkage between bundle
summarize features that are common to both displacement and transducer channel opening
sets of organs. The chapter emphasizes two may be termed nanomechanics. For now we
approaches that are of particular importance in concentrate on macromechanics.
understanding the function of an organ: 1) its Both the semicircular canals and the otoconial
biomechanics, which determines how motion of or otolith organs can be considered as inertial
the head is coupled to hair-bundle deflection; sensors. Head movements are faithfully trans­
and 2) afferent discharge, which is the result of mitted to the membranous labyrinth because it is
the several stages of transduction, and describes tethered to the skull. In the case of the semicir­
the information delivered by the vestibular nerve cular canals, the relative displacement of the
to the brain. Consistent with the overall empha­ cupula depends on the inertia of endolymph,
sis of this book, we concentrate on mammals. whose rotational motion lags behind that of the
Studies in other vertebrates have been summa­ membranous canal duct. As a result, the displace­
rized elsewhere (Lysakowski and Goldberg ment of the endolymph relative to the canal wall
2004) . is in a direction opposite that of the provoking

70
4. PHYS I OLOGY OF THE VESTI B U LAR ORGANS 71

head movement. Because of fluid continuity, the down even though the vector representing
cupula also moves backward with a volume dis­ motion rotates by 180 degrees when expressed
placement equal to that of the endolyrnph. For in earth-fixed coordinates. The motion vector,
the otolith organs, the inertial sensors are the when transformed from earth-fixed to head-fixed
otoconia, calcium carbonate crystals that give coordinates, remains constant, as does the
the upper layer of the otoconial membrane a response. For the labyrinth, then, it is best to
density greater than that of the surrounding express motion vectors in a head-fixed reference
endolymph. As a consequence of the differential frame. Yet as we shall see in Chapter 15, the
density, a linear head acceleration in one direc­ brain can combine information from otolith
tion leads to an oppositely directed shearing dis­ organs and semicircular canals to transform the
placement of the otoconial membrane relative reference frame from being head-fixed to being
to the skull. earth-fixed.
In Chapter 15, we make a distinction between
coordinate systems and reference frames. A
Rest i n g D i scharge
coordinate system is a three-dimensional vector
space in which a variable is measured. In the As first described by Lowenstein and Sand
case of the vestibular organs, the appropriate (1936), canal afferents continue to discharge
variable is head acceleration, which can be rep­ even in the absence of head rotations. A resting
resented as a vector in inertial space. At the same discharge offers several advantages. First, as
time, we can choose a reference frame in which illustrated in Figure 4. 1, it allows each fiber to
the response to a particular motion is invariant. respond bidirectionally, increasing its discharge
For the labyrinth, response is determined by when the head moves in one direction and
the orientation of head motion relative to head­ decreasing it for head motion in the opposite
fixed receptors. For example, a semicircular canal direction. A discharge increase is referred to as
will respond in the same way to a leftward head an excitation and a decrease as an inhibition.

..
rotation whether the subject is upright or upside This is so even though the decrease might

.. ... ..
,_
1 50 ,.._
�Q)
:
·o..
..:.:::

Cl)

Q)- 1 00

..::. .
...,.
Q)

. ..,,.
�-.
Ol

. ...,
:
.
(.)

.,
Cl)

.. ,
50
..
i:5 :
' -! .,
.. ,,. .
I

0
:'
0 50 1 00 1 50 200
Time, s

Figure 4 . 1 Bidirectional response of a vestibular-nerve afferent to sinusoidal


head rotations, 0.05 Hz, 80 deg/s. In its response, the discharge of this
anterior-canal afferent increases during cupular deflections away from the
canal duct and decreases during oppositely directed deflections. In passing
from excitation to inhibition, there is no discontinuity, such as would occur
were there a sensory threshold. Vertical marks indicate instants of peak excit­
atory velocity. Lower and upper horizontal lines, respectively, indicate resting
discharge before and after stimulation. (From Fernandez and Goldberg 1971.
With permission of the American Physiological Society.)
72 THE VESTI B U LAR SYSTEM

more properly be called a disfacilitation since it regularly discharging, canal afferents (Estes
is based on a decrease in transducer current, et al. 1975; Goldberg and Fernandez 1971b;
which in tum results in a decreased release of Lysakowski et al. 1995; Tomko et al. 1981b). A
excitatory neurotransmitter from hair cells (see smaller difference is observed between regular
Chapter 3). Second, resting activity can reduce, and irregular otolith afferents (Fernandez and
if not eliminate, the existence of a sensory thresh­ Goldberg 1976a; Goldberg et al. 1990a; Tomko
old. As discharge in Figure 4. 1 passes through its et al. 1981a) .
resting value, there is no sign of a discontinuity . The above results were obtained in barbiturate­
Third, resting activity provides a powerful excit­ anesthetized preparations. Similar observations
atory input to the brain, as is exemplified by the have been made in unanesthetized, decerebrate
drastic reduction in the resting discharge of preparations (Blanks and Precht 1978; Perachio
secondary neurons in the vestibular nuclei after and Correia 1983; Plotnik et al. 2005) and in
labyrinthectomy or vestibular-nerve section (see alert monkeys (Haque et al. 2004; Keller 1976;
Chapters 13 and 16). Lisberger and Pavelko 1986; Louie and Kimm
�he resting discharge is measured in the 1976; Ramachandran and Lisberger 2006;
absence of stimulation. Because semicircular­ Sadeghi et al. 2007b) . In fact, except for a modest
canal afferents respond to angular accelerations, reduction in resting discharge confined to irreg­
their resting discharge can be determined by ular afferents (Perachio and Correia 1983;
keeping the head stationary or having it move at Plotnik et al. 2005), anesthesia is found to have
a constant angular velocity. Otolith afferents little or no effect on afferent discharge proper­
respond to linear forces, including gravity. With ties in mammals, although it may result in a large
the continual presence of gravity, an absence of depression of the resting discharge in pigeons
stimulation is not possible terrestrially. But here (Anastasio et al. 1985). The modest resting
use can be made of the fact that the directional discharge reduction in mammals or the larger
properties of each otolith afferent can be approx­ effect in pigeons may reflect a direct action
imated by a polarization vector that summarizes of anesthesia on the labyrinth combined with
the response when the head is tilted in various an indirect action mediated by the efferent
directions with respect to the earth vertical vestibular system (see Chapter 5).
(Angelaki and Dickman 2000; Fernandez et al.
1972; Loe et al. 1973) . The resting (zero-force)
D i scharge Reg u larity
discharge is obtained when the polarization and
gravity vectors are orthogonal. This will occur at Some fibers have a regular spacing of action
two points during a sequence of tilts around potentials, while in other units the spacing is
pitch, roll or any other great circle (see Section irregular (Fig. 4.2). Discharge regularity has
4.3, Directional properties, for details). proved useful in classifying units (Goldberg
Resting discharge depends on species, organ, 2000) There are three reasons. First, discharge
and discharge regularity. As will be amplified in regularity is characteristic of each unit. This can
the next section, afferents can have a regular or an be seen in Figure 4.3, which plots the relation
irregular spacing of action potentials (Fig. 4.2). between the standard deviation of intervals (sd)
Background rates among regularly discharging and the mean interval ( t ) for three otolith affer­
canal afferents average 50 to 100 spikes/s, being ents whose discharge was allowed to reach a
higher in monkeys (Goldberg and Fernandez steady state at different tilt angles. The points
1971a; Lysakowski et al. 1995) than in cats for each unit, whether obtained during excita­
(Anderson et al. 1978; Estes et al. 1975; Tomko tion, rest, or inhibition, form a single relation.
et al. 1981b) or various rodents (Baird et al. 1988; Furthermore, the relations for different units do
Curthoys 1982; Hullar and Minor 1999; Hullar not intersect. Second, it is easy to quantify this
et al. 2005; Schneider and Anderson, 1976). discharge property by calculating_ the coefficient
Resting rates are lower in afferents innervating of variation (cv), the ratio of sd to t . The cv varies
otolith organs (Fernandez et al. 1972; Fernandez with t . To account for this variation, we use a
and Goldberg 1976a; Goldberg et al. 1990a) normalized statistic (cv* ) , the cv at a standard
and in irregularly discharging, as compared to mean interval. Given typical discharge rates in
4. PHYS I OLOGY OF T H E VESTI B U LAR ORGANS 73

Regular it was suggested that discharge regularity and


encoder sensitivity are mechanistically linked
(Baird et al. 1988; Ezure et al. 1983; Goldberg
et al. 1984) . Galvanic responses are much larger
in irregular (Fig. 4.4A) than in regular units
(Fig. 4.4B). There is evidence that the currents
work on afferent terminals, rather than on hair
cells or parent axons (Goldberg et al. 1984).
When galvanic sensitivity is plotted against cv* ,
a strong relation i s obtained (Fig. 4.4C). cv*
varies in the population more than 20-fold, from
less than 0.025 to more than 0.5, and there is a
associated 10-fold variation in galvanic sensitivity.
I rregu lar
A stochastic model of repetitive discharge can
be used to explain the relation between dis­
charge regularity and galvanic sensitivity (Smith
and Goldberg 1986) (Fig. 4.5). Consistent with
the treatment in Chapter 3, repetitive activity in
the model reflects the interaction of an after­
hyp erpolarization (AHP) following each spike
with synaptic and other depolarizing currents.
The random timing of synaptic quanta results in
synaptic noise that is responsible for the variabil­
ity of interspike intervals. Consider the simu­
lated train of a regularly discharging afferent
(Fig. 4.5A) . Here, each AHP is deep and slow.
50 ms
Synaptic currents are sufficiently intense so that
Figure 4.2 Discharge regularity in vestibular-nerve the mean vo1tage trajectory crosses the critical
afferents. Spike trains are shown during the resting firing level (CFL) or spike threshold. Firing is
discharge for two afferents, each innervating the ante­
rior semicircular canal in a squirrel monkey. Although
said to be deterministic. The result is a regular
both afferents have similar discharge rates, just under
discharge. To simulate an irregular discharge
100 spikes/s, they differ in the spacing of their action
(Fig. 4.5B), the AHP is made shallower and
potentials, which is regular in the top afferent and faster. Of lesser importance, there is also an
irregular in the bottom afferent. (From Goldberg and increase in the size of synaptic quanta. To reach
Fernandez 1971b. Wtth permission of the American
Physiological Society.)
the same discharge rate as shown for the regular
unit requires much less synaptic current and the
mean trajectory does not cross the CFL. Firing
is non-deterministic, occurring because of the
mammals, 15 ms provides a suitable standard occurrence of synaptic quanta (mEPSPs). As a
interval (Fig. 4.3, vertical line) . The cv* gives an result, discharge reflects the random timing of
unambiguous measure of discharge regularity, the quanta and is irregular. To see why the two
independent of discharge rate. Third, and most units also differ in their sensitivities to depolar­
importantly, fibers classified as regularly or izing inputs, we can add an externally applied
irregularly discharging differ in several other depolarizing current. In the regular unit, the
respects as well. current shifts the mean voltage trajectory slightly
Table 4. 1 summarizes some of these differ­ upward (Fig. 4.5A, thin curve) and results in a
ences for afferents, including those innervating small increase in firing rate (Fig. 4.5C). In the
the cristae and the maculae. Of the several dif­ irregular unit, a similar upward shift results in
ferences listed in the table, which are causally several previously ineffective peaks in synaptic
related to discharge regularity? Based on the noise (Fig. 4.5B, dots) now exceeding the thresh­
response to externally applied galvanic currents, old (Fig. 4.5D) and in an increase in firing rate
cv*

15
en
E • • Excitatory
en o o Resting discharge
co
C: X X I n h ibitory
(l)

_-g 1 0
0

-�
C:
0

· s;
(l)
"O
"O

.§ 5
co
C:

if)

0
I I
0 5 10 15 20
Mean interval, ms

Figure 4 .3 Quantif).ing discharge regularity. Relations between the standard deviation of


intervals and the mean interval for three otolith units differing in discharge regularity. The
points for each unit fall along a single trajectory and those for different units do not inter­
sect. Data were obtained during excitation, rest, and inhibition (see Key) . To characterize
discharge regularity, the coefficient of variation ( cv*) is calculated at a mean interval of 15
ms (vertical line) . For the three units, cv* = 0.67 (top ) , 0.20 (middle), and 0.033 (bottom).
(From Goldberg and Fernandez 1971b. With permission of the American Physiological
Society.)

..
. ,...
200
A C


.c
� 1 00 ">
:,c;
·5..
-e
"vi
C
0)

·c
Cl)
(l)
(.)
§ 0

�-0,�
Cl.) � 0. 1
ci'l
Ol co
Ol
"O
"§ 1 00 B ........ 0)

i5 . ':::!
co

z
l E
0
0 0.0 1 +-----r---r-r-r-,-,"TT"r----r---r-........-rrn-n
0 10 20 .0 1 .1
Time (s) cv*

Figure 4.4 Encoder sensitivity is higher, the more irregular is the discharge of a unit. Such
sensitivity is measured by galvanic currents delivered by way of the perilymphatic space.
Responses of an irregularly (A) and a regularly discharging unit (B) from the same animal.
Currents are delivered between 10 and 15 s (bars ) and are of the same magnitude for both
units. Cathodal currents excite(e); anodal currents inhibit (0). Responses are much larger
in the irregular unit. C. Galvanic sensitivity ,·ersus discharge regularity ( cv*) for several
units. There is a strong positive relation between the two variables. D. In these experiments,
one stimulating electrode is placed in the vestibule (white cross); the other (not shovm) is in
the middle ear. Abbreviations: aa, al, and ap, ampullae for the anterior, lateral, and poste­
rior canals; c, cochlea; cc, crus commune; s and u, sacculus and utriculus. (A-C, from Baird
et al. 1988, with permission of the American Physiological Society; D modified from
Lindeman 1969.)

74
4. PHYSI OLOGY O F TH E V ESTI B U LAR O RGANS 75

Regular

1 0 ms 66.8 _,. 6 8 .1 spikes/s


I rreg ular

66.8 _,. 92.2 spikes/s

Figure 4.5 A stochastic model of repetitive discharge simulates the discharge


of a regular unit (A) and an irregular unit (B). The t\vo units differ in their
afterhyperpolarizations (AHPs), which are deeper and slower in the regular
unit. Resting potential, 0 mV; constant threshold (Thr) for spike discharge.
The random timing of quanta (mEPSPs) introduces synaptic noise in both
cases. Quantal size and synaptic noise are larger in the irregular unit. Mean
voltage trajectory (thick curve) and the effects of a galvanic current (thin curve)
are shown for both units. The mean trajectory crosses threshold in the regular
unit but not in the irregular unit. The effects of a 1-mV depolarization are
shown in C and D for the hvo units. Sensitivity to the depolarization, measured
by the increase in discharge rate, is almost 20-fold larger for the irregular unit,
25.4 vs. 1.3 spikes/s (see numbers under voltage traces). Dots in B mark peaks
that are within 1 mV of threshold . ( Based on Smith and Goldberg 1986.)

almost 20 times greater than that seen in the obtained from a comparison of the response
regular unit. dynamics obtained with sinusoidal galvanic
The more irregular the discharge of a unit, the currents and sinusoidal head rotations (Ezure
greater is its sensitivity to synaptic or external et al. 1983; Goldberg et al. 1982). In addition,
currents. Given this conclusion we can interpret the conclusion illustrates that two discharge
the entries in Table 4. 1. Based on their greater properties-in this case discharge regularity and
encoder sensitivity, irregular fibers would be response dynamics-can be highly correlated
expected to have larger responses to sensory without being causally related.
inputs, to efferent activation, and to externally As indicated in Table 4. 1, the response dynam­
applied galvanic currents. Fiber size, while it is ics of regular and irregular afferents differ.
known to affect electrical excitability (Rushton Regular afferents have tonic response dynamics,
1951), has a much smaller effect on the galvanic similar to that expected of macromechanics.
sensitivity of vestibular-nerve afferents than does Irregular afferents are more phasic, implying
discharge regularity (Goldberg et al. 1984; that they are sensitive to the velocity of cupular
Lysakowski et al. 1995) . The only difference or otolithic displacement, as well as to the dis­
listed in the table that is not causally related to placement itself. We will return to the etiology
discharge regularity involves response dynam­ of response dynamics later ( 4.2 Semicircular
ics. Confirmation of the latter conclusion was canals, Afferent response dynamics).
76 TH E VESTI B U LAR SYSTEM

Table 4. 1 Characteristics of regularly and irregularly discharging afferents, mammalian vestibular nerve
Irregularly Discharging Regularly Discharging

Thick and medium-sized axons ending as calyx and dimorphic Medium-sized and thin axons ending as dimorphic and
terminals in the central (striolar) zone. 1 bouton terminals in the peripheral (peripheral
extrastriolar) zone.
Phasic-tonic response dynamics, including a sensitivity to the Tonic response dynamics, resembling those expected of
velocity of cupular (otolith) displacement. 2 end-organ macromechanics.
High sensitivity to angular or linear forces. (Calyx units Low sensitivity to angular or linear forces.
innervating the cristae have an irregular discharge and low
sensitivities. ) 2
Large responses to electrical stimulation of efferent fibers. 3 Small responses to electrical stimulation of efferent
fibers.
Low thresholds to short shocks and large responses to constant High thresholds and small responses to the same galvanic
galvanic currents, both delivered via the perilymphatic space.4 stimuli.
1 Goldberg and Fernandez, 1977; Yagi et al., 1977; Baird et al., 1988; Goldberg et al., 1990b; Lysakowski et al., 1995
2Goldberg and Fernandez, 1971b; Schneider and Anderson, 1976; Fernandez and Goldberg, 1976c; Curthoys, 1982; Baird et al., 1988;
Gol�berg et al., 1990a; Lysakowski et al., 1995
3Goldberg and Fernandez, 1980; McCue and Guinan, 1994; Marlinski et al., 2004
4Ezure et al., 1983; Goldberg et al., 1984; Bronte-Stewart and Lisberger, 1994

I nfo rmati on Tra n s m i s s i o n In Figure 4.6, the two units had similar
gains. There are irregular units with considerably
Information theory was devised by Shannon
higher gains, which should serve to enhance
( Shannon and Weaver 1949) to describe the
their information transmission. In fact.high­
uncertainty or entropy contained in an ensemble
gain irregular units have signal-to-noise ratios
of messages (e.g., a set of stimuli or of responses).
and Mis similar to those of regular units
The mutual information (MI) encoded by a sen­
(Hirsch et al. 2011). One might conclude that
sory system is defined as the average reduction
an irregular discharge offers no functional
in the uncertainty of the stimulus ensemble
advantage. Here, we can make two comments.
resulting from the presence of a particular
First, information theory deals with the ideal
response (Borst and Theunissen 1999; Rieke
handling of information transmission, not the
et al. 1997). Of particular interest in the present
actual way this occurs. Second, the calculation
context is the question as to how efficiently regu­
emphasiz�s signals (gains) and noise (discharge
lar and irregular spike trains encode features of
irregularity). Regular and irregular units differ
the stimulus. For continuous ensembles, MI is
in several other respects. It would seem safe
related to the signal-to-noise ratio characterizing
to assume that irregular units play distinctive
the stimulus or response. Interspike-interval
roles in the processing of vestibular sensation.
variability during identical stimulus conditions
To deduce this role, we need to understand
may be viewed as noise. From this, it might be
the contributions of both kinds of afferents to
expected that irregular units would not be as
central processing, a topic we will consider in
efficient in stimulus encoding. This conclusion
Chapter 7.
can be illustrated by a linear-systems analysis
that deduces the optimal filter that, when con­
volved ,vith the spike train, provides a recon­
structed stimulus with minimal error from the 4.2 S EM I C I RC U LAR CANALS
actual stimulus (Borst and Theunissen 1999;
Rieke et al. 1997). Figure 4.6 compares a regular Canals and otolith organs differ in their func­
and an irregular unit (Sadeghi et al. 2007a). The tions: the former sense angular rotations, while
regular unit is much better at estimating the the latter monitor linear motions and head tilts.
actual stimulus. Corollaries of this result are that Consistent with this conclusion, canal afferents
the regular unit is better at transmitting informa­ respond to angular head rotations. In acute
tion about the stimulus and at detecting small preparations these same afferents can also
changes in head motion. respond to linear forces (Estes et al. 1975;
.

f
so
A
Regular afferent - Head velocity
gi 40

CF= 0.53 Stimulus reconstruction j


M l = 0.48 bits/spk
0

'5o -
iI:
-400 0 400
Time (ms)
50

-- �
- en
Q)
> ---
Cl
Q) [
"O "O
ro -
1 0
1 50
2
ro -
g' o.. [
·c .!!:,,,
U:: 1 00

IIIIIIII IIIIIIIIIIIIIIIIIIIIIIIIIIIIII IIIIIIIIII IIII IIII IIIIIIIII I I IIIIIIIIIIIIIIIIIIIIIIIIIIIIIIII


1 00 ms

rl±
B

I rregular afferent
C F= 0. 1 3
M l=0.08 bits/spk
-400 0 400
c Time (ms)
:B (/)
Q) ---
> Cl
50

Q)
"O "O [
ro -
1 0

��
� ,..j 50

I IIIII IIII II I IIIIIII I IIIIIIIIIIII IIIIII IIIIII IIIII Il I I IIIII IllIIll Illllll II II IIIIIIIII IIIIIll
C Cl.
·c
[

U:: 1 00

1 00 ms

Figure 4 . 6 Stimulus reconstructions for two horizontal semicircular canal


units to rotational broadband (O to 20 Hz) Gaussian pseudorandom stimula­
tion. Time-dependent firing rates (lower traces) are shown for a regular (A)
and an irregular unit (B). For each unit, an impulse response (see insets) , when
convolved with the spike train, gave a reconstructed stimulus that minimized
the error between the reconstruction and the actual stimulus (upper traces ).
The regular unit provided a more faithful reconstruction. CF, coding fraction;
MI, mutual information. (From Sadeghi et al. 2007a. With permission of the
Society for Neuroscience.)

77
78 THE VEST! B U LAR SYSTEM

Goldberg and Fernandez 1975; Perachio and (Lindeman 1969; Lowenstein and Wersall 1954).
Correia 1983). One interpretation of the linear­ The canals are arranged in coplanar pairs with
force responses is that they are artifacts caused the two horizontal canals forming a pair, as do
by thermal or other gradients introduced into the anterior canal on one side and the contralat­
the endolymphatic ring by the acute exposure of eral posterior canal (see Fig. 2.3). Based on
the temporal bone (Goldberg and Fernandez the particular canals (A, H, P) and the side of
1975). Consistent with this hypothesis, such the head (L, R), the three pairs are referred to
responses are not found when less invasive as LHRH, LARP, and RALP. Any head rotation
recordings are made in chronic animals (Correia causing excitation or inhibition from a canal
et al. 1992; Somps et al. 1994). will result in an oppositely directed response
from the contralateral coplanar canal. It is
the difference in discharge between coplanar
D i rect i o n a l Properties
canals that is interpreted by the brain as a head
From physical principles, it would be expected rotation.
that !n response to head rotations the displace­
ment of endolymph in each canal duct would be
Macro m echan i cs and t h e
proportional to the cosine of the angle between
To r s i o n - Pe n d u l u m Model
an optimal plane and the plane of motion. Were
fluid flow in each of the three canals indepen­ Early workers (Breuer 1874, 1875; Crum
dent of one another, the optimal plane should Brown 187 4; Mach 187 4) speculated that the
correspond to the geometric canal plane. But semicircular canals, because of their toroidal
because each canal has two openings into the topology, were involved in sensing rotational
utriculus, fluid flows in the three canals are not head movements. The nerve endings in the canal
independent and the optimal planes can deviate were presumed to be sensitive to fluid pressure
from the geometric planes (Rabbitt 1999). (Mach 1874) or fluid motion (Crum Brown 1874;
Detailed anatomical measurements of canal Breuer 1874). There was a difficulty ,vith these
planes are available for several species (Blanks theories. Pressure or motion in an unoccluded
et al. 1972, 1985; Curthoys et al. 1975; Reisine canal should outlast head movements by only a
et al. 1988). Experimentally, the optimal planes, few milliseconds. Yet, it ,,,as known that vestibu­
measured from afferent discharge, are within 10 lar reflexes and sensations can persist for many
degrees of the geometric planes (Estes et al. seconds following certain motion profiles, for
1975; Haque et al. 2004; Reisine et al. 1988). example when a subject on rotating platform
Ewald (1892) studied directional properties is suddenly stopped after a period of constant
by applying or removing pressure from canal angular velocity. Two explanations could be
ducts while monitoring head and eye move­ offered for the apparent discrepancy between
ments. He summarized his finding in two laws: fluid motion and the central manifestations of
responses were in the plane of the stimulated canal excitation: (1) biomechanical models of the
canal (First Law) and one stimulus direction canals were deficient or (2) central mechanisms
(push or pull) led to distinctly larger responses in could result in a perseveration of the activity
each canal (Second Law). Afferent recordings triggered by afferent input. As it turns out,
(Lowenstein and Sand 1940a) paralleled Ewald's both peripheral and central mechanisms are
findings. All afferents innervating a given canal involved in prolonging vestibular responses. The
have the same directional properties. Deflections central mechanisms are termed "velocity stor­
of the cupula toward the utriculus are excitatory age" and will be considered later (Chapters 9
for the horizontal canal and inhibitory for the and 11). Here, we concentrate on peripheral
vertical canals. Excitatory directions are those mechanisms.
that led to larger responses in Ewald's experi­ Classical theories ignored the influence of the
ments, and they are correlated with hair-bundle cupula on fluid motion. A reason for this miscon­
morphological polarization, which is uniform ception was the shrinkage of the cupula during
in each crista (see Fig. 2.9C), but oppositely histological fixation. To appreciate the role of
directed in the horizontal and vertical cristae the cupula, it had to be visualized in situ vvithout
4. PHYS I OLOGY O F THE VESTI B U LA R ORGANS 79

fixatives. This was done by Steinhausen (1931), length, Len· The entire centerline, including the
who injected dye into the unfixed membranous ampulla and the utriculus, has a length, L.
labyrinth of fish. He discovered that the cupula, To deduce the model's behavior, we consider
rather than sitting atop the crista like a cocked the forces acting on dl, an infinitesimal segment
hat, extended across the lumen from the top of of the endolymph in the canal duct (Fig. 4. 7B).
the crista to the vault of the ampulla, where The theory leads to the second-order equation
· J, dl J, dl
pQ· 'f - + 8nµQ· 'f � + kQ
it formed a hydraulic seal. In response to rota­
L A (l) L A (l)
tions, the cupula was observed to move across
the ampullary vault like a "swinging gate." CD CD
(4.1)
Importantly, when the cupula was displaced, it = -p � ax (l)dl
was observed to take several seconds to return to L

its neutral position. Dahlman (1935) confirmed where Q is the volume displacement of the
Steinhausen's observations. Later studies found endolymph, the line integrals are calculated over
that the cupula, rather than moving like a "s,ving­ the length of the canal duct, and a /Z) = Ra/l),
ing gate," is tethered both to the vault of where a /Z) is the linear acceleration equivalent
the ampulla and to the crista and moves like a to a/l), the component of head angular accel­
diaphragm constrained along its entire circum­ eration in the canal plane. To denote �erivatives,
ference (Hillman and McLaren 1979). we use the dot convention, Q = dQ I dt
Steinhausen (1931) not only provided evi­ and Q = d 2 Q I dt 2 . Because of fluid continuity, Q
dence for the role of the cupula but also is the same across any cross-section, including
developed a quantitative model of the macro­ the cupula, in which case Q = Acvx cv = A cui:Xc ur
mechanics. The model is formally similar to a Here, x cv and x cuP are the linear displacements
torsion pendulum, such as illustrated in text­ of the canal duct (CD) and the cupula (C UP),
books of elementary physics. In the model, exci­ respectively; A cv and Ac uP are the corresponding
tation is assumed to be proportional to the cross-sectional areas. A(l) is much smaller in the
displacement of the cupula-endolymph relative canal duct than in the cupula and the utriculus. As
to the canal wall. The original model, as well as a result, the first two terms on the left side need to
subsequent versions (Jones and Spells 1963; be evaluated ' only over the length of the canal
Ramprashad et al. 1984; van Egmond et al. duct, L ev · (A derivation of the equation can be
1949), treated the canal duct similarly, but dif­ found in an appendix,"Derivation of the Torsion­
fered in the way they handled the enlarged vol­ Pendulum Model" at < http://www.oup.com/us/
umes of the ampulla and utricular sac. The first thevestibularsystem> .) The equivalent lumped
model to deal satisfactorily with this complex · model is depicted in Figure 4. 7C.
geometry treated the mechanics in a piecewise We will use Equation 4 . 1 to deduce the
manner (Oman et aL 1987). response of the cupula to head rotations. But we
As in other versions of the torsion-pendulum first consider why the canals do not respond to
model, the piecewise model leads to a second­ linear forces. When there is an angular accelera­
order differential equation that relates the tion, there is a net couple, -pLRa cos0 that
angular acceleration of the head (a) to the dis­ moves the endolymph. Now consider a linear
placement of the endolymph in the canal duct acceleration, ax, that is constant in magnitude
(x cv) . Three factors determine the relation: (1) and direction. Here, the couples at points 180
the geometry of the canal, in this case its radius degrees apart will cancel. As a result, the net
of curvature, R(l), and its cross-sectional area, couple will vanish and there will be no movement
A(l), both expressed as functions of l, the posi­ of the cupula and endolymph. While cancellation
tion along the streamline (Fig. 4.7A); (2) the does not depend on the exact geometry of the
physical properties of the endolymph, including canal, it does depend on the density of the endo­
its density, p, and viscosity, µ; and (3) the elastic­ lymph being constant along its entire length (L).
ity of the cupula (k). To simplify the geometry, A variable density provides a basis for the
we assume that the centerline of the canal is a caloric response in which hot or cold water is
circle of radius, R, and that the canal duct (CD) flushed into the ear canal and changes the den­
has a constant cross-sectional area, A CD ' and a sity of endolymph in the external-most part of
80 THE VESTI B U LAR SYSTEM

acose

p y

-i]-_____.
-
C
( Viscous drag)
dF0

____.

_.

A(X)= ax(C)+ x(C)
dm

dFc
( Cupula elasticity)

Figure 4 . 7 A. Planar section of the membranous semicircular canal. R, radius


of curvature; r, cross-sectional radius of canal duct. Streamline has a length, L.
(Modified from Curthoys and Oman 1986.) B. Free-body diagram of an infini­
tesimal section of the canal duct of length, dl, and cross-sectional area, A. Q,
volume flow of endolymph. Fluid pressure, P. dFD ' force on fluid due to vis­
cous drag. a, angular acceleration of the head in a plane tilted at an angle, 8,
from the effective canal plane. The component of linear acceleration of the
head in the canal plane is Ra cos 8. (Modified from Rabbitt et al. 2004a) C.
A lumped model of macromechanics. As the head is accelerated _!n space,
a.x (.e) , endoiymph accelerates backv1ard relative to the canal wall, x(.e). The
backward movement is opposed by two restoring forces, the viscous drag
exerted by the canal wall (dFD) and the elasticity of the cupula (dFc).

the horizontal-canal duct. Placing a subject in a becomes sensitive to linear forces during caloric
supine position allows gravity to act in the plane stimulation (Young and Anderson 1974). While
of the horizontal canal. ·with warm water in the the caloric response was recognized in the 19th
ear canal, endolymph in the canal duct becomes century (Goltz 1870), Barany (1906) is credited
less dense and will rise under the influence with the realization that at least part of the
of gravity. The result is a buoyant force that response is due to the buoyancy of endolymph.
moves the cupula towards the utriculus and, There are also non-buoyancy components (Coats
hence, is excitatory. Cold water produces and Smith 1967; Minor and Goldberg 1990;
the opposite reaction: cupular displacement Paige 1985), as was convincingly demonstrated
is towards the canal duct and is inhibitory. by the persistence of a caloric response during
Afferent recordings confirm that the canal spaceflight (Scherer et al. 1986). Non-buoyancy
4. PHYS IOLOGY O F TH E VESTI B U LA R ORGANS 81

t-
components could include direct temperature where x i s average cupular displacement. Here,
effects on various stages of transduction and/or fA = Ac iAc ur' fL = ULCD , and the subscripts
temperature-dependent localized expansions ( CD and CUP) refer, respectively, to the canal duct
and contractions of labyrinthine fluids, which and the cupula.
could result in a cupular displacement ( Scherer Equation 4.5 can be solved by standard
and Clarke 1985). We ,vill come back to this methods (see, for example, Edwards and Penney
topic in Chapter 16. 2000). To illustrate the workings of the equation,
Returning to rotational stimulation, we note we show solutions for two head movements. The
that Equation 4. 1 is a second-order, linear first is a brief motion that might occur during a
ordinary differential equation of the form voluntary head saccade intended to bring a target
Mx + Bx + Kx = - f (t) located in the visual periphery into the center of
(4.2)
gaze (Fig. 4.8A). Here, the magnitude of cupular
The behavior of the equation is determined position almost exactly parallels angular head
by two time constants. As we shall see, velocity. In the second example, angular velocity
B 2 >> 4MK , in which case the system is is gradually built up and then the motion is
m:'erdamped and the time constants can be abruptly stopped (Fig. 4.8B). As velocity grows,
evaluated as cupular displacement increases in an exponen­
r1 = B l K tial manner, its magnitude initially paralleling
8nµLcv angular velocity, but eventually reaching an
= --- (4 . 3 )
asymptote proportional to angular acceleration.
kA�D
When the motion is stopped, the cupula is flung
and
in the opposite direction and only slowly returns
r2 = M / B to rest. While the stimulus in Figure 4.8A resem­
pACD (4.4 ) bles a typical, everyday voluntary head move­
Snµ ment, the motion in Figure 4.8B is atypically
with 't 1>>'t2 Substituting these values of the time slow and the terminal deceleration is atypical in
constants into Equation 4.2 gives not being matched by an immediately preced­
ing, oppositely directed acceleration of similar
magnitude. Interest in the latter motion stems in

B 2 C
2

- Cupula position
- Head velocity
0 -2

0.0 0.5 0 20 40 0 20 40
Time, s

Figu re 4 . 8 Solution of the torsion-pendulum equation (thick dots in A and


thick lines in B and C) for three angular head-velocity profiles (thin lines). In
all calculations, long time-constant, --c 1 = 5 s. A A brief rotation consisting of the
excitatory half-cycle of a 2-Hz sine wave. The response magnitude parallels the
rotation profile. B. A velocity ramp (acceleration step) lasting 20 s, followed
by a sudden stop. The response consists of an exponential build-up, a large
displacement in the opposite direction, followed by an exponential return
to the baseline. C. A velocity step leads to a transient exponential response
after an initial step. In all cases, cupular deflection was inverted to facilitate
comparison with head velocity.
82 TH E VEST! B U LAR SYSTEM

part from its historical use in clinical vestibular angular head velocity and ,.vill be in conflict with
testing and in part because it mimics a play­ information provided by vision and propriocep­
ground maneuver (see Chapter 1.4). While a tion. This is an example of the conclusion, stated
voluntary head movement cannot accomplish in Chapter 1, that vertigo arises when there is a
the motion, it can be obtained by whole-body conflict or mismatch between different sensory
spins on a turntable or by self-generated spin­ channels. In contrast, the motion depicted in
ning. As the reader was asked to verify in Chapter Figure 4.8A is unaccompanied by such symp­
1, the brief deceleration terminating the motion toms, reflecting the fact that cupular displace­
can lead to vertigo and postural instability. Later ment is congruent with head velocity and
in this chapter, we will see that afferent dis­ non-vestibular inputs.
charge more or less parallels cupular displace­ The relation between head movements and
ment and that the brain interprets the discharge cupular displacement can be summarized by a
as proportional to angular head velocity. During so-called Bode plot, in which the gain and phase
the later parts of the rotation and particularly in of cupular displacement is plotted as a function
t�e post-rotatory period, cupular displacement of sinusoidal frequency (Fig. 4.9). In the figure,
and afferent discharge no longer indicate true vertical dashed lines are drawn at the two

· co
C

CJ 0 . 1

B 90

0)
Q)
"O

$ 0 -+-------;.------==,_-.::,-------�----­
cu
(/)

..c
a..

Frequency, Hz
0.001 0.Q1 0.1 10 1 00 1 000

Figure 4 . 9 Bode plot of torsion-pendulum model, including gain (A) and


phase (B) re angular head velocity. Corner frequencies, 0.025 and 25 Hz.
Vertical dashed lines divide the spectrum into low-, mid-, and high-frequency
regions in which the output parallels angular acceleration, velocity, and
displacement, respectively.
4. PHYS I OLOGY OF T H E VESTI B U LAR ORGANS 83

so-called corner frequencies, J;_ = 1 / 2nr1 and cupular displacement. Psychophysical studies
f2 = 1 / 2nr2 , related to the time constants of indicate a similar parallel when the sensation of
Equation 4.5. The lines divide the frequency turning is estimated by the subject ( Guedry
axis into a low-frequency region (f < f ) , a mid­ 1974). From both its macromechanics and
frequency region (betweenJi andi), and a high­ the way the brain interprets its signals, the semi­
frequency region (f > t). In the low-frequency circular canals function as sensors of angular
region, the system is an acceleration transducer: velocity for midband frequencies.
gain re velocity increases in proportion to f, The restriction to midband frequencies is
which is equivalent to a constant gain re accel­ important since at very low frequencies the
eration, and the phase lead re velocity approaches canals serve as angular-acceleration transducers.
90 degrees, in register with acceleration. The This makes physical sense since only head accel­
system approximates a velocity transducer in erations result in the inertial forces that produce
the mid-frequency range: gain re velocity is fluid motion. It is the reactive forces resulting
nearly constant and phase re velocity is near from fluid motion, including endolymph viscos­
0 degrees. In the high-frequency region, the ity and cupular elasticity, that determine stimu­
•. system becomes a displacement transducer: gain lus specificity. These ideas can be illustrated by
re velocity is inversely proportional to frequency the response of the cupula-endolymph to a pro­
and phase re velocity approaches a 90-degree longed step of angular velocity (see Fig. 4.8C).
lag, which places response in phase with head There is a response at the beginning of the step,
displacement. These three regimes reflect the which decays exponentially with a time constant
three terms on the left side of Equation 4.5 cor­ of about 5 s. Unlike the situation for a constant
responding to the influence of endolymph mass head acceleration (see Fig. 4.8B), there is no
(Mx) , endolymph viscosity (Bx ) and cupular steady-state response, which is why we do not
elasticity (Kx). As sinusoidal frequency is low­ sense the rotation of the earth about its axis or
ered from high values, the system is dominated around the sun.
first by the mass term, then by the viscous term, The torsion-pendulum model can be used to
and finally by the elastic term. The right-hand compute cupular displacements for the dynamic
side of the equation remains proportional to range over which the semicircular canals oper­
angular head acceleration (a). A solution in ate. Based on studies in humans, the threshold
terms of cupular displacement (x) requires a for detection of long-duration head accelerations
double integration of a when the mass term is near 0. 1 deg/s2 (Clark 1967; Guedry 1974).
dominates, a single integration when the viscous From voluntary head movements in humans, a
term dominates, and no integration when the large, typically encountered rotation can be
elastic term dominates. taken as 400 deg/s (Armand and Minor 2001;
The mid-frequency region or midband extends Liao et al. 2005). From the torsion-pendulum
over three decades. Typical voluntary head equation, we can compute the average displace­
movements fall comfortably within the midband ment of the cupula for long-duration accelera­
(Armand and Minor 2001; Liao et al. 2005), as tions (Oman and Young 1972). By assuming that
do the head perturbations accompanying loco­ the cupula adopts a parabolic shape, we can cal­
motion (Grossman et al. 1988) and other daily culate the displacement immediately above the
activities ( MacDougall and Moore 2005). Most crista, at a distance corresponding to the average
of the energy in such head movements falls in a height of the tallest stereocilia. In the center of
frequency band from 0.5 Hz to possibly 20 Hz, the crista, the height is 5 to 10 µm, whereas at
with a peak around 2 to 5 Hz. For such head the periphery of the neuroepithelium, hair bun­
movements, the semicircular canals function as dles can be much longer (LeWis et al. 1985; Lim
angular-velocity transducers. To see how the 1976). The calculated displacement of central
brain interprets neural signals emanating from stereocilia at threshold is 1.5 to 3 nm, about
the canals, we can consider the eye move­ 1 % to 2% the diameter of an individual stereo­
ments produced during vestibular stimulation. cilium (Hunter-Duvar and Hinajosa 1984). For
As will be considered in Chapter 9, it is the a 400 deg/s midband rotation, the calculated
velocity of eye movements that parallels central hair-bundle displacement is 0.6 µm,
84 THE VESTI B U LAR SYSTEM

within the dynamic range for mechanoelectric dynamics to the slower movements of large ani­
transducer (MET) currents in some vestibular mals. To determine whether sensitivity changes
hair cells (Holt et al. 1997). At threshold, the pres­ in the manner suggested by Melvill Jones (Jones
sure across the cupula is 1 x 10-3 dyne/cm2 . and Spells 1963; Melvill Jones 1974), we can
The numbers emphasize both the sensitivity compare for several species experimental values,
and the versatility of the organ. Calculated obtained from afferent recordings, with values
threshold displacements and pressures are simi­ predicted from canal geometry (Fig. 4. 10) (see
lar in magnitude to those obtained in the mam­ Yang and Hullar 2007). In all cases, gains were
malian cochlea at hearing threshold (0Ison 2001; confined to regular units. Except possibly for
Robles and Ruggero 2001). At the other end of the cat, there is reasonable agreement between
the amplitude scale, evidence from afferent empirical gains and the predictions of the Jones­
recordings indicates that the cupular or hair­ Spells theory.
bundle displacements occurring during large While the agreement is reassuring, some of
head rotations do not challenge the structural the assumptions implicit in the theory require
integrity of the transduction machinery. comment. First, it was tacitly assumed that any
particular canal has a single sensitivity. In mam­
mals, there is an approximately 10-fold variation
I nterspecies Var i at i o n s a n d
Canal Dimensions
Several of the parameters appearing i n the 0.5
0
torsion-pendulum model (Equation 4.5) are
related to the dimensions of the semicircular T
canal. So, for example, average cupular displace­ J.
ment, x, re angular head velocity, m, for mid­ '"i"U)
band head rotations is oi
Q)
"O

• *
8µ V
x= ( 4 .6)
■ ■
• Mouse
where R is the radius of curvature and r is the · Rat
A Guinea Pig
cross-sectional radius of the canal duct. At a con­ T Gerbil
stant temperature, the physical constants, p and Rabbit

l
µ, are fixed and the ratios, �'\. and f1 , are nearly □
0 Squi rrel monkey

!::. Rhesus monkey


Chinchilla
constant across species, so, to a first approxima­
tion, sensitivity (x I m) , should be proportional 0.0 V Cat

to the product, Rr2. 0.0 2.5 X 1 0-4

Relations between torsion-pendulum param­


Figu re 4. 1 0 Relation between the mean rotational
eters and canal geometry (van Egmond et al.
gain of regular semicircular canal afferents ( cv*
1949) were exploited by Melvill Jones (Jones and
Spells 1963; Melvill Jones 1974), who compared < 0.10) and the dimensions of the membranous
the value of Rr2 across species. In mammals, canal. R, radius of curvature; r, cross-sectional
radius of the membranous canal duct. Line is best­ J
fitting regression passing through the origin.
Rr2 increased 20-fold over a 105 range of
Morphological data, mouse ( Calabrese and Hullar
body weights. Melvill Jones speculated that the
2006; A Lysakowski, personal communication); rat,
increase in sensitivity with body mass reflected
an adaptive matching of canal sensitivity to guinea pig (Curthoys and Oman 1986); gerbil, rabbit,
movement repertoire. Furthermore, it was sug­ chinchilla (Ramprashad et al. 1984); squirrel monkey,
cat (Igarashi 1967); rhesus monkey (Jones and Spells
1963). Physiological data, mouse (Yang and Hullar
gested that the frequencies defining the mid­

2007); rat, guinea pig (Curthoys 1982); gerbil


band should move dmvnward with body mass. In
( Schneider and Anderson 1976); rabbit (Stahl 1992);
general, large animals, being more sluggish than
small animals, would benefit from an increase squirrel monkey (Lysakowski et al. 1995); chinchilla
in canal sensitivity. The downward shift of the ( Baird et al. 1988); rhesus monkey (Haque et al. 2004);
midband was interpreted as matching canal cat (Tomko et al. 1981b).
4. PHYSIOLOGY OF T H E VESTI B U LAR ORGANS 85

in afferent sensitivity systematically related to nonlinear distortions being typically less than
discharge regularity (Baird et al. 1988; Curthoys 10%. Much of the nonlinear distortion reflects
1982; Goldberg and Fernandez 1971b; Haque asymmetries bet\;veen excitatory and inhibitory
et al. 2004; Lysakowski et al. 1995; Ramachandran responses . Third, gains and phases at a given
and Lisberger 2006; Sadeghi et al. 2007b; frequency are almost constant as peak velocity
Schneider and Anderson 1976; Tomko et al. varies over a vvide range (Fig. 4. 11 inset).
1981b). An even larger, almost 100-fold varia­ Gains and phases for typical regular and
tion in afferent sensitivity is seen in lower verte­ irregular afferents are compared in Figure 4. 12
brates (Boyle and Highstein 1990; Brichta and with those expected from macromechanics. Data
Goldberg 2000a; Honrubia et al. 1989). Much of are from the squirrel monkey (Fernandez
the gain vaiiation seen in mammals can be and Goldberg 1971; Goldberg and Fernandez
ascribed to differences in encoder sensitivity 1971b). Throughout a frequency range extend­
(see Section 4. 1, Discharge regularity), implying ing up to 0.5 Hz, the gains and phases of the
that macromechanics is not a limiting consider­ regular afferent are close to those expected
ation. Second, Melvill Jones suggested that from the torsion-pendulum model, but above
variations in time constants could be used to 0.5 Hz, there are discrepancies. The irregu­
match canal dynamics with the dynamics of larly discharging afferent shows deviations from
head motions . But the bandwidth of typical expected values at both low and high frequencies.
head movements is much narrower than the Consider the high-frequency deviations . As
canal midband, in which case the hypothesized frequency is increased above 0.5 Hz, there are
dynamic matching would be of marginal adap­ progressive phase leads and gain enhancements
tive value. Furthermore and contrary to predic­ not predicted by the torsion-pendulum model.
tions, when closely related animals are compared, Deviations are larger in the irregular unit and
canal dimensions tend to be larger in the more can be interpreted as indicating that afferent
agile species (reviewed in Spoor 1998) . Third, response, r(t), is proportional to a weighted sum
the variation in the dimensions of the vestibular of cupular displacement, x(t) , and velocity, dx/dt,
end organs may reflect constraints other than with the relative weight of the two factors vary­
those related to the presumed need to match ing across units, but not ,vi.th frequency. That is,
afferent signals to head movements-for exam­ afferent response
ple, the need to fit the labyrinth into different­
r(t) oc x(t) + rv dx ! dt (4.7)
sized skulls or to innervate the vestibular nuclei

l
whose size grows with body mass (Matano 1986; In the formulation, x(t) is the expected
Stephan et al. 1981). response based on the torsion-pendulum model
and 'tv is a fixed weight characteristic of each
afferent. The larger the weight, the more con­
Affe rent Response Dyn a m i cs
spicuous is the velocity term. Equation 4. 7
Do dynamics parallel the torsion-pendulum describes the high-frequency deviations in both
model or are they modified by any of the several units of Figure 4. 12. In the irregular unit,
stages of vestibular transduction . interposed 'rv = 0.080 s provides a best fit to the data. Even

I
between macromechanics and afferent-nerve though the discrepancy is smaller in the regular
discharge? To study the question, we can exam­ unit, it can be fit by Equation 4. 7 with a value
ine afferent responses to sinusoidal head rota­ of 'rv = 0.015 s. That the differences in response
tions. But first we need to determine whether dynamics of afferents can be described by the
the system behaves linearly as would be consis­ variation of one or a few parameters in a trans­
tent vvith the linear differential equation defin­ fer function may be termed parameterization .
ing the torsion-pendulum model (Equation 4. 1). A practical consequence of parameterization is
As illustrated in Figure 4. 11, responses resem­ that differences between units can usually be
ble those of a linear system in three ways. First, characterized by determining their gains and
there is no evidence of a threshold or disconti­ phases at a single frequency. Parameterization
nuity as the curves pass through zero response. holds for afferents recorded in mammals other
Second, responses are close to sinusoidal with than the squirrel monkey, even though the
....
86 THE V ESTI B U LAR SYSTEM

1 00

50 •
■ ■
■ ■ ■ ■
■ ■ 0
�a.,
■ ■ 280
1 40



Head velocity, deg/s
..:.:: ■ ■
·5.. • ■
CJ)
a.,­
CJ)
C
0 0
0..
CJ)
a.,

■ ■ •

-50 •

• 256 deg/s
■ 1 28
0 0 64
A A 32

90 0 -90 -1 80 -270
Phase re head velocity, deg

Figure 4.11 Steady-state sinusoidal responses of the same anterior-canal unit


as in Figure 4.1 to 0.05-Hz sinusoidal head rotations, different peak angular
velocities. Responses are sinusoidal in form and do not show a discontinuity on
passing from excitation to inhibition. Zero degrees, peak excitatory velocity.
Response leads stimulus. Inset plots gain and phase re head velocity as func­
tions of peak velocity. Both gains and phases are almost constant even though
stimulus magnitude changes 16-fold (inset). (From Fernandez and Goldberg
1971. With permission of the American Physiological Society.)

high-frequency velocity sensitivity may involve declines and post-acceleratory secondary


fractional operators (Baird et al. 1988; Curthoys responses, including an undershoot in rate fol­
1982; Schneider and Anderson 1976; Tomko lowing excitation and an overshoot following inhi­
et al. 1981b), rather than the integral operator bition. Adaptation is also reflected as phase leads
used in Equation 4. 7. in the response to very low-frequency (::;0.025
We now consider the low-frequency deviation. Hz) sinusoidal head rotations (Fernandez and
This is best seen in the responses to long-duration Goldberg 1971) (see Fig. 4. 12) .
angular-velocity trapezoids, which are shown for What mechanisms contribute to differences
a regular unit (Fig. 4. 13A) and an irregular unit between afferents in their response dynamics?
(Fig. 4. 13B). Except for an asymmetry between Conceivably, any of the transduction steps
its excitatory and inhibitory responses, the regular interposed between· head motion and afferent
unit conforms to expectations from the torsion­ discharge may contribute to afferent diversity.
pendulum model. During long-duration constant Two complementary strategies have been used
angular accelerations and decelerations, excit­ to identify potential mechanisms: ( 1) electric
atory and inhibitory responses build up exponen­ currents have been used to byp ass various trans­
tially with time constants near 5s. The return to duction stages and (2) recordings have been
the resting discharge from either kind of response made from hair cells and afferent terminals to
is also exponential. In contrast, the discharge of distinguish contributions at various stages of the
the irregular unit shows per-acceleratory response trandusduction chain.
4. PHYSIOLOGY O F T H E VESTI B U LA R ORGA N S 87

,.
.75 l rreg
• - -•
- 1 00 Reg
0-- · - · -0

.50
0) Torsion
Q) �

;i,
-o -80 \ Pend u l u m
c
· c3 .25 · c3
I
Q)
0
Q)

• /6
e /

I
0
� -60
_,.,,,., C1l
Q)
0- -o
- � .00
C


..c

0)

0)
0 � -40
_J C1l
0

iO • ..c
0..
-.25

-20
0

-.50

-.75

.0 1 25 .050 .250 1 .00 4.00 . 0 1 25 . 050 . 250 1 .00 4.00


.025 . 1 00 .500 2.00 8.00 .025 . 1 00 . 500 2.00 8.00
Frequency ( H z)

Figure 4 . 1 2 Response dynamics of semicircular-canal afferents in the squirrel


monkey. Because their responses are nearly linear, the afferents can be char­
acterized by their gains (left) and phase leads (right) re angular head velocity
in response to sinusoidal head rotations. Such Bode plots are compared for an
irregular (Irreg) unit, a regular (Reg) unit, and the torsion-pendulum model.
(From Goldberg and Fernandez 19716. With permission of the American
Physiological Society .)

Cupular Deflection deduced from afferent discharge may be consider­


ably higher than that determined by actual obser­
While there may be a differential displace­
vation of cupular motion (Rabbitt et al. 2009).
ment between the center and planar edges of
the cupula, the effect is likely to be small and
confined to relatively high frequencies (Highstein
Hair- Bundle Micromechanics
et al. 2005; Rabbitt et al. 2004a) . The reasoning
is confirmed in mammals; here there is a con­ Hair bundles are shorter in central/striolar
centric organization of the crista such that regu­ zones than in peripheral/extrastriolar zones
larly discharging afferents with indistinguishable (Lewis et al. 1985). A looser or viscous coupling
response dynamics are found throughout the between the shorter bundles and the o,·erlying
peripheral zone, including the base of the crista accessory structure might contribute to the more
along its longitudinal length and the apex near phasic response dynamics of central/striolar
the planum semilunatum (Baird et al. 1988) . afferents. Here, the evidence is contradictory.
This observation suggests that regional variations Mechanical stimulation of semicircular canals
in cupular mechanics are unlikely to make a has been compared with electrical polarization
major contribution to the diversity of response of the endolymphatic space in the toadfish
dynamics seen in mammals. ( Highstein et al. 1996) . Presumably, hair-bundle
It has usually been assumed that cupular dyna­ deflections are involved in the responses to
mics can be inferred from the low-frequency mechanical but not to galvanic stimulation. The
behavior of regular units. A recent study in the two modes of stimulation had different effects
toadfish suggests that the slow time constant on the most tonic and most phasic afferents,
88 THE VEST! B U LAR SYSTEM

A
1 20

� 0
Q)

B
0
15
U)
1 60

1 20

60

0
0 40
Time, s

T3

T1 TS

Figu re 4 . 1 3 Some afferents show a low-frequency adaptation, which can be


demonstrated with velocity trapezoids, in this case consisting of two 40-s veloc­
ity ramps (acceleration, 7.5 deg!s2) separated by a 60-s (A) or a 90-s (B) velocity
plateau. An irregular unit (B) shows adaptation, which consists of per-stimulus
response declines and post-stimulus secondary responses. A regular unit (A)
shows little adaptation as its response increases exponentially during velocity
ramps and decreases exponentially during the velocity plateau. There is a 30-s
break in A so that the inhibitory ramps for the hvo units are in register. The
velocity profile is shown below and consists of a rest period (Tl), an excitatory
velocity ramp (T2), a velocity plateau (T3), an inhibitory velocity ramp (T4),
and a final stationary period (T5). (From Goldberg and Fernandez 197la.)

consistent with a micromechanical contribution recordings with sharp electrodes may distort
to response dyn amic diversity. hair-cell physiology.
Hair-cell recordings of receptor potentials in
the toadfish crista during mechanical stimulation
Transducer Adaptation
have led to a different conclusion (Highstein
et al. 1996). Here, sinusoidal stimulation resulted The decline in transducer currents during sus­
in relatively flat (tonic) response dyn amics. tained hair-bundle deflections is an obvious can­
Furthermore, there was relatively little adapta­ didate for shaping response dynamics. Yet, here
tion to step stimulation. Taken at face value, the too results are not entirely consistent. Work has
results would imply either that the hair cells been done in the utricular macula where striolar
sampled only supplied tonic afferents or else afferents have more phasic response dynamics
that micromechanics do not contribute to than extrastriolar afferents. Recordings in early
dynamic diversity. But as the authors themselves postnatal mice did not show any differences in
point out (Highstein et al. 1996), the results adaptive properties between striolar and extrast­
should be interpreted cautiously. In particular, riolar hair cells (Vollrath and Eatock 2003).
4. PHYS I OLOGY O F TH E V ESTI B U LAR ORGANS 89

In contrast, striking differences were seen in or Comparisons for a population of afferents are
near the striola of the frog utriculus among hair seen in Figure 4. 14B. There is a strong correla­
cells differing in bundle morphology (Baird 1994) . tion between spike and quanta! phases. At the
same time the discrepancy between the two
phases grows, the larger are the phase leads or,
Basolateral Currents
equivalently, the more phasic the response
Such currents, by shaping receptor potentials, dynamics. About two thirds of the total variation
could contribute to the response dynamics of i� spike phase can be ascribed to quanta! phase
hair cells. Despite the large repertoire of such and is likely, therefore, to be caused by presyn­
currents (Eatock and Lysakowski, 2006; Guth aptic factors. The data do not allow us to specify
et al. 1998), it is unlikely that they contribute to the presynaptic factors involved, which may
the diversity of afferent response dynamics. The include any of the several mechanisms described
negative conclusion is based on comparisons above. One mechanism considered in Chapter 3
between two populations of hair cells in the is vesicle turnover. It would be expected that
turtle posterior crista, those near the center there would be a depletion of vesicles during
(torus) and those near the edge of the organ excitation and their replenishment during inhi­
(planum). Despite large differences in afferent bition. Both effects would lead to a phase
response dynamics (Brichta and Goldberg advance. Such processes may underlie adapta­
2000a), the two sets of hair cells were indistin­ tion in the fish sacculus (Furukawa and Matsuura
guishable in the gains and phases of their cur­ 1978; Furukawa et al. 1982) and in chick
rent-clamp responses to sinusoidal currents ( Spassova et al. 2004) and mammalian auditory
( Goldberg and Brichta 2002b). It should be afferents ( Goutman and Glowatzki 2007).
noted that the results were obtained in enzymat­
ically dissociated hair cells and might be differ­
Postsynaptic Mechanisms
ent in vivo .
As illustrated in Figure 4. 14B, about one third
of the phase variation across units cannot be
Quanta/ Release
accounted for by presynaptic mechanisms. This
As we saw in Chapter 3, neurotransmitter is fraction may be due to any of several postsynap­
released from hair cells in quanta! packets. The tic actions. The first concerns postsynaptic neu­
sinusoidal modulation of quanta! rate presum­ rotransmitter action. As summarized in Chapter
ably reflects presynaptic mechanisms. By com­ 3, neurotransmission from hair cells to afferents
paring quanta! release to spike discharge, we can involves the release of glutamate acting on
estimate the hair-cell (presynaptic) and afferent­ AMPA receptors. Receptor activation occurs
terminal (postsynaptic) contribution to response too fast (;:::, 1 ms) to contribute to response dynam­
dynamics. This has been done in the turtle pos­ ics (Dingledine et al. 1999). Processes having
te1ior crista ( Holt et al. 2006b). Figure 4. 14A, slower kinetics include receptor desensitization,
based on the response of a single unit to canal­ clearance by transporters, and non-quanta!
duct indentation at 0.3 Hz, compares variations transmission. These have only minimal effects
in spike rate and quantal rate over a single cycle. on response dynamics, as is indicated by the fact
The stimulus has an excitatory peak at 270 that there is no consistent discrepancy between
degrees (vertical bar). Spike activity leads the quanta! phases (Fig. 4. 14B) and depolarization
mechanical stimulus by 66 degrees (red curve), phases (Fig. 4. i4C).
about 35 degrees more phase advanced than The discrepancy between postsynaptic depo­
quanta! activity (blue curve). Were there com­ larization and spi�te discharge can be considered
plete agreement between the spike and quantal a form of postsynaptic adaptation. Traditionally,
phases, this would suggest that the spike phase is such adaptation was ascribed to continual spik­
entirely determined by presyn aptic (hair-cell) ing activity ( Goldberg and Fernandez 1971a;
mechanisms. A postsynaptic contribution is sug­ Taglietti et al. 1977). To study this, sinusoidal
gested by the 35-degree discrepancy between galvanic currents were introduced via the peri­
the two curves. lymphatic space in mammals and were shown to
90 TH E VEST! B U LAR SYSTEM

40
g>
90
B g> 90 C
,
A

..
5000 "O "O

.. . ... .
(I)
0 (I)
' i.
� 60 � 60
,,
C
Ill ..c ..c
� a. a.
� C , ' ,'
0 ,'
OJ � 30 � 30 ,'
,
N
,,
_ro
-'-
-�
co
YI cu ,, '
0 0 a o
0
0
g- 0

0 1 80 360 0 30 60 90 0 30 60 90
Phase, deg Spike phase, deg

Figure 4.1 4 Comparing phases of quantal activity, postsynaptic depolariza­


tion, and spike activity for turtle posterior-crista units in response to a 0.3-Hz
sinusoidal stimulus, an indentation of the posterior canal duct. Spike activity
was collected before and synaptic activity after spikes were blocked by TTX.
Approximately 30 degrees of the 60-degree variation in spike phase can be
accounted for by quantal phase, which is interpreted as a presynaptic contribu­
tion; the remaining 30 degrees may be the result of postsynaptic factors. A.
Phase histograms of spike activity (left ordinate) and of quantal rate (right ordi­
nate) versus stimulus phase for an individual unit, turtle posterior crista. Peak
excitatory stimulus, 270 degrees (vertical bar). Phase re stimulus of spike rate
( 63 degrees) leads that of quantal depolarization (48 degrees). B. Phases of
quantal rate (ordinate) versus that of spike discharge (abscissa) for 23 units.
Dashed line, unity diagonal. Solid line, best-fitting straight line. C. Phases of
postsynaptic depolarization versus that of spike discharge, same 23 units. There
is no significant difference between the two phase measures of synaptic activ­
ity, quantal rate (B) and postsynaptic depolarization (C). (Modified from Holt
et al. 20066.)

directly activate the postsynaptic spike encoder for 2-Hz sinusoidal head rotations are plotted
(Ezure et al. 1983; Goldberg et al. 1982) . The versus cv* for several extracellularly recorded
currents resulted in sinusoidal responses with units in the chinchilla. The gain plot provides
phase leads of 10 to 15 degrees that were similar evidence for two populations. Units in the first
in all units. Because of the similarity, this form of population range from very regular (cv* = . 025 (to
adaptation could not contribute to afferent modestly irregular (cv* :::::; 0.25) and gain increases
diversity. In the toadfish, some of the phase lead linearly with cv*. For the irregular units of the
observed in irregularly discharging, phasic affer­ first group, gains approach 2 spikes•s-1/deg•s-1 .
ents could be blocked by GABA-B antagonists The second population consists of the most irreg­
(Holstein et al. 2004b) . This was related to the ular units in the sample; the gains of these affer­
observation that some hair cells are GABAergic ents are about five times lower than would be
(Holstein et al.2004a) . The mechanisms by predicted from an extrapolation of the regression
means of which GABA might exert its presumed line for the first group (Fig. 4. 15A) . Remarkably,
postsynaptic actions remains to be elucidated there is only a single relation between phase
(Holt et al. 2006b) . and cv* for the units of the two populations
(Fig. 4.15B) . Phases range from lags of 5 to
10 degrees in the most regular units to leads near
Variat i o n s i n Gai n and Phase
30 degrees in the most irregular units. The pres­
Gain and phase vary with discharge regularity ence of two groups, first established in the chin­
(Baird et al. 1988; Haque et al. 2004; Hullar et al. chilla (Baird et al. 1988; Hullar et al. 2005), was
2005; Ramachandran and Lisberger 2006; Sadeghi later confirmed in monkeys (Haque et al. 2004)
et al. 2007b). In Figure 4. 15A,B, gain and phase and the mouse (Yang and Hullar 2007).
/

4. PHYS IOLOGY O F T H E VESTI B U LAR O RGANS 91


:
1 .0 1 .0 :
A C

.. .-.:: ' .
·u0
ent 1

Q)

-· . .
-0
Cll �
Q)


0
.c
0

.
Peripheral · . • . • •
� -:·)°· •
N .
8
I / :
••.· • •
C
'cil •
('.)

0
.1 . 1 +--.....-'"....-....-,-.-.-m----.-.--.--.-,�
.01 .1 .01 .1

60 60
B D
� •
0
'ti
40 40
Q)

Cll
Q)


.c
20 20

N
8
I 0 0

Q)
•o Calyx
a.. -20
.c -20 Dimorphic
xBouton
. U nlabeled
· -40 +----.--�...,....,.""T'"T""T'T"T----r-.----.--.-......--r-r,
-4 0
.01 .1 .01 .1
Coefficient o f variation (cv*)

Figure 4 . 1 5 Gains and phases versus normalized coefficient of variation (cv* ).


Semicircular-canal afferents in the chinchilla responding to 2-Hz sinusoidal
head rotations. Each point represents one unit. A, B. Extracellularly recorded
units. Based on their gains, units fall into two groups. Straight line in A is best­
fitting power law relation between gain and cv* for one of the groups. Straight
line in B is best-fitting semilogarithmic relation between phase and cv* for all
units. C, D. Corresponding data for intracellularly labeled units (see Key).
Unlabeled units are shown as small dots. Labeled units: 14 calyx units,
26 dimorphic units, and one bouton unit. Calyx units are the most irregularly
discharging afferents and have the largest phase leads and distinctively low
gains. Gain and phase of dimorphic units increase with cv*; regular units are
found in the peripheral zone and irregular units in the central zone. The one
bouton unit was regular, had a low gain and phase, and was located in the
peripheral zone. Gains, spikes•s· 1/deg•s· 1 ; phases in degrees re head velocity
with positive values indicating phase leads. (From Baird et al. 1988. With per­
mission of the American Physiological Society.)

Two factors contribute to the gain versus cv* phasic response dynamics of irregular units.
relation for the first group. By far the more When the influence of response dynamics is
important of these is the sensitivity of the post­ eliminated, the gain curve of the first group par­
synaptic spike encoder, which can be measured allels the relation between galvanic sensitivity
by the galvanic sensitivity of individual afferents and cv* (see Fig. 4.4B). Since galvanic sensitivity
(Goldberg et al. 1984; Smith and Goldberg provides a measure of encoder gain, the parallel
1986). The other factor is the high-frequency between the two relations implies that the syn­
gain enhancement associated with the more aptic input to the encoder is nearly constant for
92 THE VESTI B U LAR SYSTEM

units of the first group. By the same reasoning, reflects their small axon diameter, which makes
since their galvanic sensitivity is in line with their them difficult to impale. Fortunately, their small
discharge regularity, synaptic input is about 5x diameter also makes it possible to identify bouton
lower for units of the second group. units in extracellular recordings by their small
conduction velocities. This was done in the
squirrel monkey (Lysakowski et al. 1995). As
Affe rent Morphology and Phys i o l ogy
might be expected, the presumed bouton units
The first attempt to relate afferent diversity are regularly discharging with small gains and
with morphology compared physiological esti­ near-zero phases.
mates of fiber size with discharge regularity The results, particularly those for dimorphic
and other discharge properties ( Goldberg and units, emphasize the importance of regional
Fernandez 1977; Lysakowski et al. 1995; Yagi variations in discharge properties. Within this
et al. 1977). A more direct approach, done in the single morphological class, there is an almost
chinchilla, involves the intra-axonal labeling of 10-fold variation in rotational gain between
physiologically characterized fibers (Baird et al. afferents innervating the peripheral and central
1988) . With this method, fibers are impaled and zones, a comparable variation in encoder
their physiology is characterized, after which a (galvanic) sensitivity, and a large variation in
label is iontophoresed from the recording micro­ response dynamics as judged by a 30- to
electrode and travels down the axon to its peri­ 40-degree difference in the 2-Hz rotational
pheral terminations. Similar studies have been phases of afferents innervating the two zones.
done in non-mammalian vertebrates (Boyle et al. Reinforcing the importance of regional varia­
1991; Brichta and Goldberg 2000a; Myers tion are the similarities in discharge properties
and Lewis 1990; reviewed in Lysakowski and between peripheral dimorphic and bouton
Goldberg 2004). units and between central dimorphic and calyx
The results of intracellular labeling in the fibers.
chinchilla are summarized in Figure 4. 15C,D. Of the various differences in afferent dis­
Unlabeled (small symbols) and labeled fibers charge, only the large discrepancies in rotational
(large symbols) are included in the figure. As gains of calyx fibers and their central dimorphic
expected from extracellular labeling studies (see neighbors do not reflect regional variations. It
Fig. 2. 14), intracellularly labeled calyx units were should be emphasized that the lower sensitivity
found to terminate in the central zone. These of calyx fibers is unrelated to their encoder
fibers are irregularly discharging units with rela­ sensitivity. In fact, calyx units have the most
tively small gains and large phase leads ( •). In irregular discharge, the largest phase leads, and
short, calyx units are the second group of irregu­ the largest galvanic responses of the entire
larly discharging units identified in extracellular population. Their relatively large phase leads
recordings. Dimorphic units belong to the first suggest that the rotational gains of calyx units
group ( o). The physiology of dimorphic units should increase with frequency faster than
depends on their locations in the crista. Those those of irregular dimorphic units. This is the
terminating in the central zone are irregularly case at least in the chinchilla. As frequency
discharging with large gains and phases. In addi­ increases towards 10 Hz, the gains of calyx units
tion to their lower gains, calyx units are slightly approach those of irregular dimorphs (Hullar
more irregular in their discharge and have et al. 2005). On the other hand, gain versus
slightly larger phases than central dimorphs. frequency curves for calyx and high-gain irregu­
Peripheral dimorphs are regularly discharging lar units in the monkey parallel one another
and have small gains and phases. (Ramachandran and Lisberger 2006; Sadeghi
Only one intra-axonally labeled bouton unit et al. 2007b).
could be traced to its neuroepithelial termina­ The higher synaptic gains of dimorphic as
tion in the peripheral zone. Like peripheral compared to calyx units might be explained by
dimorphic units, the bouton unit was regularly the auxiliary inputs the former receive from type
discharging and had small gains and phases. The II hair cells. A scheme in which type I and type
paucity of labeled bouton fibers presumably II inputs sum linearly leads to an estimate that
4. PHYSIOLOGY OF T H E VESTI B U LAR ORGANS 93

each type I hair cell contributes three times the Dyn a m i c Ran ge of Affe rent D i scharge
synaptic input to an afferent than it receives
In a preceding section (see Fig. 4. 10), we saw
from each of its bouton contacts (Baird et al.
that afferent responses behaved linearly at
1988). A 3: 1 ratio is less than the 10: 1 ratio in the
least for moderate angular head velocities. But
number of ribbon synapses made with individual
response should eventually reach an upper limit.
calyx and bouton endings (Lysakowski and
The question arises as to the value of the upper
Goldberg 1997). The discrepancy in these ratios
limit and the head velocities at which such satu­
may be related to the distinctive features of
rating nonlinearities become evident. Excitatory
synaptic transmission between type I hair cells
responses as a function of head velocity are shown
and their calyx endings. As reviewed in a previ­
for three groups of afferents in Figure 4.16
ous chapter (Section 3.5, Synaptic transmission
(Plotnik et al. 1999). Relations behveen excit­
involving type I hair cells), synaptic transmission
atory response (r) and head velocity (v) can be fit
may be less efficient at calyx than at bouston
by the empirical function
endings. Another possibility is that synaptic
transmission involving calyx endings may differ rMAX v
r(v ) = (4.8)
for the two afferent groups. V1 1 2 +V

A Regular units B Irregular dimorphic units C I rregular calyx units


;
500 ; 500
;
;
;

i� ;
;
;


;
·5.. ;

Q)
(/)
C
0
Cl..
(/)
Q)

0 0
0 500 1 000 0 500 1 000 0 500 1 000
Head velocity (deg/s)

1 000 400
c7i
Cl)
I
Ill
if)
Q) 200 Cl..
_:,,:


<
·5.. Cl)


0
(')
Q) 0
(/) 0
C
ci:
(/) -200 S
a: �
Q)
f/2..
-1 000 -400
0.0 0. 4
Time, s

Figure 4 . 1 6 Relations between response and angular head velocity for three
populations of semicircular-canal afferents in the chinchilla. Responses and
head velocities, average values during last 0.5 s of a 2-s period of constant
angular acceleration. For each population, a non-linear function (solid line,
Equation 4.11) based on the average parameters for the population is com­
pared with the corresponding linear function (dashed line) . Regular units, cv*
< 0.05 (A); non-calyx units, cv* > 0.20 (B); calyx units ( C). Distinction between
calyx and non-calyx units based on a quadratic discriminant function. D.
Calculations of the responses in the squirrel monkey to a head saccade, peak
ve�ocity of 400 deg!s and peak acceleration of 10,000 degls2. In the calculation
we used the mean velocity and acceleration sensitivities from Lysakowski et al.
( 1995). ( Data in A-C from Plotnik et al. 1999.)
94 THE VESTI B U LA R SYSTEM

which is recognized as the Michaelis-Ment�n If, as suggested, the role of these afferents is
equation of enzyme kinetics. Here rMAX is the related to their relatively low rotational gains
maximum response and v112 is a constant. For and phasic response dynamics, it is unclear why
v < < v 112 , response is linear with a gain or slope these discharge properties would require the
of rM Jv112 • As velocity increases beyond v112 , elaborate structure of the calyx ending.
response approaches an asymptote, rMAX ' For all
units, rMAX is 300 to 500 spikes/s (Plotnik et al.
1999). 4.3 OTOLITH ORGANS
Is this rMAX adequate to handle the head rota­
tions encountered in everyday life? The angular Afferents innervating otolith organs respond
velocities obtaining during locomotor activities to linear but not to rotational forces acting on the
seldom exceed 200 degls (Grossman et al. 1988). head (Goldberg and FBrnandez 1975). Linear
Even rapid voluntary head movements are rarely forces arise when the head is accelerated along
larger than 400 degls, although accelerations can straight lines. Because of the presence of gravity,
be near 10,000 degls2 (Armand and Minor 2001; afferents also signal head tilts ,�nth respect to the
Liao et al. 2005). In Figure 4. 16D, we have earth vertical. In fact, because the otolith organs
calculated the response to a 400 deg/s head sac­ are linear-force sensors, their response should
cade assuming linear rate-intensity relations. be indistinguishable for a head tilt and a linear
The peak response of calyx units (300 spikes/s) is acceleration of appropriate magnitude and direc­
close to rMAX' while that of irregular dimorphs tion (Angelaki et al. 2004; Fernandez and
( 1, 100 spikes/s) clearly exceeds this limit. Goldberg 1976b).
These calculations suggest that the low gains
of calyx units allow them to encode rapid head
Directional Properties
movements without reaching excitatory satura­
tion. But a low gain cannot be the entire story The responses of otolith afferents depend on
since regular units have even lower gains and the orientation of linear forces with respect to
smaller peak responses. This suggests that the the head. Directional properties for each fiber
distinctive function of calyx units involves the can be summarized by a unit polarization vector,
combination of its ability to handle fast head v = (x, y, z), expressed in head-fixed coordinates
movements and the other properties that distin­ (see Fig. 1. 1) 1 and reflecting the polarization
guish irregularly discharging units. Of the other vectors of the innervated hair cells (Angelaki and
properties, phasic response dyn amics would Dickman 2000; Fernandez et al. 1972; Loe et al.
seem most relevant since they allow the peak 1973; Tomko et al. 1981a). The directional prop­
responses of calyx and other irregular afferents erties of peripheral otolith neurons can be termed
to anticipate peak head velocity and, thus, could one-dimensional (ID) as they can be character­
compensate for delays in various reflex pathways ized by a single vector. In contrast, many
( Huterer and Cullen 2002; Minor et al. 1999; central otolith neurons have two-dimensional
Ramachandran and Lisberger 2005). (2D) tuning summarized by a response ellipse
Type I hair cells and calyx endings are a recent (Angelaki et al. 1993; Angelaki and Dickman
evolutionary feature, being seen only in reptiles, 2000; Bush et al. 1993) . Two-dimensional tuning
birds, and mammals (Lysakowski 1996; Wersall can result from the convergence of otolith inputs
1956) . These animals are distinguished from differing in both their spatial and temporal
their aquatic and amphibian ancestors in having properties (Angelaki 1992) (s,ee Chapter 7).
mobile necks, allowing for head movements on a
nearly stationary body. It can be supposed that
the low-gain calyx afferents, seen in turtles as
well as mammals, were an adaptation allowing 1. The coordinate system illustrated in Figure 1 . 1 has been
the semicircular canals to accommodate the generally adopted in the literature and is used here. Some
studies of otolith afferent discharge (Fernandez et al. 1972;
rapid rotations made possible by the freeing of Fernandez and Goldberg 1976a-c; Loe et al. 1973) used
head movements from the rest of the body. This a different system. We have transformed the data to be
kind of reasoning raises an intriguing question. consistent with Figure 1. 1.
4. PHYSI OLOGY OF T H E VESTI B U LA R ORGANS 95

The lD spatial tuning of peripheral otolith rate. For the unit illustrated in Figure 4. 17,
neurons is illustrated by the response of a utricu­ s = 28 spikes/s•g and d0 = 40 spikes/s. v = (0.707,
lar unit to pitch and roll head tilts (Fig. 4. 17A). 0.707,0.035) is halfway between the vectors
As the head is tilted to various positions, there is pointing out the nose and out the ipsilateral (left)
a sinusoidal modulation of discharge rate about ear. Its small z component implies that v lies
an average or zero-force discharge (d/ A sinu­ close to the horizontal plane, which is taken as
soidal modulation implies that the response for the plane of the horizontal semicircular canal.
any tilt position is proportional to the component The assumption that response (d- d) is lin­
of the polarization vector coincident with the early related to effective force (F) is examined in
gravity vector. To see this, we use matrix algebra, Figure 4. 17B, C. It is seen that the relation is
specifically a rotation matrix, to transform the nonlinear with excitatory responses being larger
gravity vector, g = (0 , 0 , - 1 ) in earth-fixed coor­ than inhibitory responses. The nonlinearity
dinates, to g = (sin P, sin R cos P, - cos R cos P) in becomes more obvious when centrifugal forces
head-fixed coordinates; R is the roll angle and P of more than 1 g are introduced (Fig. 4. 1 7B).
is the pitch angle. The force (F) acting on otolith Remarkably, points for centrifugal forces ( •)
receptors is calculated as the scalar product (F) and static tilts ( O) overlap even though they
of the transformed gravity vector ( g) and a head­ represent different physical circumstances. As
fixed polarization vector, (v = x,y,z), i.e., noted previously, the equivalence between grav­
ity and a linear acceleration is to be expected of
F = v•g
a linear-force sensor. To handle the nonlinearity
= x sin P + y sin R cos P - z cos R cos P (4.9) we calculate F by linear algebra (Equation 4.9)
Presumably, v reflects the morphological
and consider the response to be a nonlinear
function of F. In the particular case of Figure
polarization of the hair cells innervated by the
4. 16C, response can be fit by the second-order
afferent. We tentatively assume that the dis­
polynomial, d - d0 = s(F + aF 2 ) with s = 32.5
charge, d (R, P) = sF + d 0 , is a linear function of
spikes/s•g and a = 0.41. The use of a quadratic
F. The sensitivity factor (s in spikes/s•g) converts
nonlinearity holds only for forces in the range
F to a discharge rate. d0 is a resting (zero-force)
±1 g. A more adequate formulation is presented
discharge. Here and elsewhere, g is the accelera­
2 later (Fig. 4.24).
tion of gravity, 980 cm/s . Equation 4.9 becomes
A vector depiction of directional properties
F = y sin R - z cos R (4.9a) requires that forces orthogonal to a unit's polar­
ization vector should be ineffective. Tests of this
for pure rolls (P = 0) and prediction are shown in Figure 4. 18 for a unit
F = x sin P - z cos P (4.9b) whose polarization vector lies near the horizon­
tal plane of the utricular macula. Centrifugal
for pure pitches (R = 0). Since F and d(R, P) are forces were used to determine the effects of
sinusoidal functions of tilt angle for pure rolls or orthogonal forces. Two kinds of orthogonal forces
pure pitches, a Fourier analysis can be used to have to be considered: compressional forces
extract the coordinates of V = sv = (X, Y, Z) . directed perpendicular to the macular plane and
Sensitivity (s ), the response obtained when shearing forces in the macular plane. Compres­
the polarization and gravity vectors coincide, sional forces are ineffective. Not only do they
is estimated from the length of V, (i.e., not lead to a response (Fig. 4. 18C), but they also
s = -Jx 2 + Y 2 + Z 2 ). The normalized polariza­ do not influence responses to simultaneously
tion vector, v = Vis, summarizes directional applied shearing forces (Fig. 4. 18E). Contrary to
properties. To calculate d0, we take the discharge predictions, orthogonal shearing forces lead to
rate obtained when the transformed gravity excitatory responses ( Fig. 4 . 18B). When response
vector, g , is orthogonal to v (i.e., F = 0). The is plotted versus force direction, excitatory
condition is met at two points during pitches, responses are produced by shearing forces span­
rolls, or any other tilt series around a great circle. ning 225 degrees, not the expected 180 degrees
The two points are recognized as being 180 (Fig. 4. 18D). The last result is inconsistent with
degrees apart, yet having the same discharge a vector treatment of directional selectivity.
96 THE VESTI B U LAR SYSTEM

80 A
80 C

40
60

en �Q)
� �
·o.
en ·o.
0 en
en
0

• •• •••
(I) (I)

Tilt Angle (deg)


-1 80 40
a.
C
1 80
en


0


a:
Q)
B


en
0
.c

• •
80 20

40 0
J3

/
X

0 -20
0 0
Time (s) Force (xlg)
10 20 -1 .23 1 .23

Figure 4.1 7 Responses of an otolith afferent to static tilts and to centrifugal


forces. A. Responses to static rolls (e e) and pitches (0 0). Lines are best
sinusoidal fits to roll (-) and pitch (---) responses. Maximum excitatory
responses, ipsilateral rolls and forward pitches. B. Responses to centrifugal
forces, parallel (excitatory, e) and anti-parallel (inhibitory, 0) to polarization
vector, the latter determined from static tilts in A. Plateau force, 1.23 g.
Excitatory response is larger than inhibitory response. C. Comparing responses
to static tilts and centrifugal force. Non-linear relation between response and
force is the same for centrifugal forces (•) and static tilts ( 0). Fit is a quadratic
polynomial. (From Fernandez and Goldberg 1976a. With permission of the
American Physiological Society.)

Three comments can be made. First, the effect hair bundle by the stretching of filamentous
cannot be explained by a convergence of hair­ strands, including tip links, connecting the
cell inputs differing in their direction properties kinocilium and stereocilia in adjacent ranks. The
(Fernandez et al. 1972). Second, the orthogonal­ result should be a fanning of the hair bundle and
shear effect is sufficiently small that a vector an opening of transducer channels for either
treatment still provides a useful approximation. direction of orthogonal shear. As expected from
Third, the effect is consistent with microme­ hair-bundle geometry, responses during orthogo­
chanics. Only the kinocilium and possibly the nal shears are smaller than those produced by
tallest stereocilia are directly attached to the shears parallel to the polarization vector.
otoconial membrane (Kachar et al. 1990; Lim Inhomogeneities in the overlying otoconial mem­
1984). This being the case, an orthogonal shear brane have been proposed as contributing to
will be communicated to the remainder of the orthogonal shear responses (Kondrachuk 2002).
. -,...·-· ·""· ..
4. PHYS IO LOGY O F TH E V ESTI B U LA R O RGANS 97

90 A D

.
40

.
• •
• •

60 20

. ..
• •
-� . .

• •
°
ss 224.5
0 •

,...
o
o:axsmo�mr
0 ,�
o�•
0

i.:s:.
Q)
30 o
o�go� �
i.:s:. •
·o..

,.._,.,.,_. ,,__.
Q)

·o..
-20
Q)
0)
--, .!:!;,

cii
Q) -1 .80
YAW (degrees)
Cl)
0 1 .80
60 B C
()
..c.

0
0..
0
Cl) E
Cl)
Q)

0

� ..Jo&o#61:B�• ��
40

30
--, 20 •
0

C •
90 0

0 X

,...
��-�..,.--,�

���· -20
0
• •

0
0
0

60
I I --, I
20 40 60 -1 .23 0 1 .23
Time (sec) Force (xlg)
0

Figure 4.1 8 Comparison of responses to shearing and compressional forces


for a utricular afferent. Polarization vector pointed out the nose. Responses are
to centrifugal-force trapezoids, 1.23 g peak force. A. Responses to shears paral­
lel (excitatory, e) and anti-parallel (inhibitory, 0) to the unit's polarization axis.
B. Shears orthogonal to polarization axis, force through ipsilateral (e) and con­
tralateral ( 0) ears, lead to excitatory responses, ::::25% of maximal excitatory
response to parallel shears. C. There are no responses when the forces are
orthogonal compressions exerting pressure ( •) or traction ( 0) on the macula.
D. Response versus angle between polarization and shear force vectors.
Excitation is seen for angles subtending 225 degrees. E. Responses to parallel
shears of varying magnitude when gravity exerts pressure ( •) or traction ( 0)
on the macula. Not only do compressional forces not lead to responses, they
also do not alter the responses to simultaneous shearing forces. (From
Fernandez and Goldberg 19766. With permission of the American Physiological
Society.)

Otolith afferents are sensitive to shearing forces, part of the utricular macula lies in a horizontal
but not to compressional forces. This finding has plane, while its anterior part curves upward.
implications for the complementary functions of Utricular afferents, at least those from its main
the utricular and saccular maculae (Fernandez and part, should have horizontally disposed polariza­
Goldberg 1976a; Tomko et al. 1981a). The main tion vectors (see Figs. 2.5 and 2.6). Responses to
98 THE VEST! BU LAR SYSTEM

static tilts are shown in Figure 4. 19A-H for eight than in the supine (180 degrees) positions. Once
units independently assigned to the superior again, discharge is seldom abolished in any head
vestibular nerve (SN). We suppose that such position.
units arise from the utricular macula, although One way to appreciate the difference in direc­
some of them may supply the anterior part of the tional properties for the two populations is to
saccular macula (see Fig. 2.2). For all eight units, compare the angles their polarization vectors
there is little difference in discharge rates for make with the horizontal (utricular) plane.
prone (O degrees) and supine (180 degrees) posi­ Angles are less than 30 degrees for most utricu­
tions, consistent with their polarization vectors lar afferents and more than 60 degrees for most
having relatively small z components and, thus, saccular afferents (Fig. 4. 19J). Another way is to
lying near the horizontal plane. Some units are plot normalized vectors for the two maculae in
excited by ipsilateral rolls (Fig. 4. 19A- E), others spherical coordinates. The utricular vectors are
by contralateral rolls (Fig. 4. 19F-H), some by best viewed from the +z axis, where they are
downward pitches (Fig. 4. 19D,E H), and others seen to be within 30 degrees of the horizontal
by upward pitches (Fig. 4. 19A,B,F) . Some units plane (Fig. 4.20, left, outer ring). There is a pre­
respond almost equally to pitches and rolls ponderance of +y vectors, but otherwise the vec­
(Fig. 4. 19B,D,F), others predominantly to rolls tors are widely distributed within this ring.
(Fig. 4. 19C,G). Stated in terms of vector coordi­ Saccular vectors are best viewed from the ipsilat­
nates, there are units with various combinations eral ear, where they are seen to align near the
of ±x and ±y coordinates. This diversity is consis­ vertical axis (Fig. 4.20, right). The results are
tent \Vith the fanlike disposition of morphologi­ consistent with the conclusion that the utricular
cal polarization vectors in the utricular macula macula provides a broad, two-dimensional rep­
(Fig. 4.20, left bottom) . In fact, each of the units resentation of linear forces acting in the horizon­
can be assigned an approximate macular location tal plane, while the saccular macula adds the
based on its vector components (Fig. 4. 19L). third or vertical dimension. During locomotion
Neither the sensitivity (s) nor the resting dis­ and many other everyday activities, vertical
charge (d0 ) is correlated with the directional linear accelerations are much larger than their
properties of utricular afferents. Typical values horizontal counterparts (MacDougall and Moore
are 40 to 90 spikes/s for d0 and 20 to 50 spikes•s- 2005; Pozzo et al. 1990), indicating the impor­
1/g for s. In only a few units and none in Figure tance of the saccular macula during these behav­
4. 19 is discharge abolished at any tilt angle. iors. Utricular afferents are most sensitive to
Quite different directional properties are seen small tilts from upright, suggesting that the latter
in units recorded from the inferior vestibular organ may be especially involved in the mainte­
nerve (IN), which innervates the main part of nance of an upright posture.
the saccular macula (Fernandez et al. 1972; In the squirrel monkey, utricular units excited
Fernandez and Goldberg 1976a; Tomko et al. by ipsilateral roll tilts outnumber those excited
1981a) (see Fig. 2.2). Such units usually have by contralateral roll tilts by a 3 : 1 ratio (Fernandez
polarization vectors with large z components et al. 1972; Fernandez and Goldberg 1976a). A
(Fig. 4. 191). Based on the sign of the component, similar preponderance of ipsilaterally excited
saccular units have maximum rates near the utricular units was observed in the cat (Loe et al.
supine (+z) or prone (-z) positions. From the 1973). In contrast, in the chinchilla, ipsilaterally
polarization map (Fig. 4.20, right bottom) , the excited utricular units were only in a slight major­
units should be located dorsal (+z) or ventral (-z) ity (Goldberg et aL 1990a) . The ratios based on
to the line of polarization reversal. As is illus­ tilt responses can be compared to the ratio of
trated in Figure 4. 191, saccular units with +z and utricular areas whose polarization vectors point
-z components differ considerably in their rest­ laterally and medially. In both the squirrel
ing values. Higher values of d0 are found in +z monkey and chinchilla, the proportion of ipsilat­
units than in -z units. Sensitivities are slightly erally excited units is slightly larger than would
higher for +z units. One result of these differ­ be predicted on an areal basis. For the monkey
ences is that the discharge rates of +z and -z saccular macula, there were approximately equal
units are much closer in the prone (O degrees) numbers of units with +z and -z components
...,... Roll
A 0--1) Pitch s
1 20
o
80 40 �
l -X, -Y

40 -X,+y
0

G

0
B
1 20
1 20

80
80

40

�Cl)
en 40

·a.
U)
0 �
C 0
"@ 80 1 20 H
Cl)
OJ

i
(.)
40 80
(/)

0 +Y 40 ��
+X, -Y
so l D
0

>
a. �
0

40
1 20
x
+�

: i:.�£¾
80

40
'o,. $>• • .; '
o

O +X, +y 0 -z
,----,---,---,----,-----,--,---,----,

-1 80 -90 0 90 1 80
-1 80 -90 0 90 1 80
Roll Contra Roll lpsi

� __,..
Roll Contra Roll l psi
Pitch Up Pitch Down
Pitch U p Pitch Down
..,.__
J L
ant

� 10 ed

0 30 60 post
Angle From Horizontal Plane

Figu re 4.1 9 Discharge rate from several otolith afferents in the squirrel monkey are plotted as functions of tilt
angle about pitch and roll axes. A-H. Each graph represents data from one unit recorded from the superior
vestibular nerve (SN) and presumably innervating the utricular macula. I. Two units recorded from the inferior
vestibular nerve (IN) and presumably innervating the saccular macula. J. Static-tilt data, such as sho�rn in A-1,
were used to calculate a polarization vector for each unit. Vectors for SN (gray-filled bars) and IN units (unfilled
bars ) differ in the angle they make with the horizontal plane. SN vectors lie near the horizontal (utricular)
plane, while IN units lie near the vertical (saccular) axis. K. Axes. L. Locations of utricular afferents (A-H)
inferred by comparing polarization vectors calculated from tilt data with morphological-polarization maps (see
Fig. 2.9). (From Fernandez and Goldberg 1976a. With permission of the American Physiological Society.)

99
1 00 THE VESTI B U LAR SYSTEM

Utricular Saccular
X z

Fig u re 4.2 0 Distribution of polarization vectors, otolith afferents, squirrel


monkey. Each point represents the polarization vector for a single unit in
spherical coordinates. Units were assigned to the utricular or saccular maculae
based on the semicircular-canal units encountered in the same puncture. •
and X, vectors pointing, respectively, towards and away from the reader. Left:
Utricular vectors are best viewed from the top of the head. The further a point
is from the z pole (center of plot), the larger is the horizontal component of its
vector. Most points for utricular vectors lie in the outer ring, signifying large
horizontal components. Right: Saccular vectors are best viewed from the ipsi­
lateral (left) ear. Most saccular vectors have large z components and lie along
. I

an axis slightly tipped with respect to the vertical. The circled points were
obtained from the inferior vestibular nerve after section of the superior ves­
tibular nerve .. Below each spherical plot is the polarization map for the corre­
sponding macula. Directions: A, anterior; P, posterior; M, medial; L, lateral; D,
dorsal; V, ventral. Coordinates follow the same convention as in Figure 4.19K.
(From Fernandez and Goldberg 1976a. With permission of the American
Physiological Society.)

(Fernandez and Goldberg 1976a). This is so 60 µm high, covering the underlying neuroepi­
despite the fact that the area dorsal to the thelium of the utricular or saccular macula and
striola is considerably larger in the monkey than made up of three layers (Fig. 4.21) (Kachar et al.
its ventral counterpart (Lysakowski and Goldberg 2000; Lim, 1984). Calcium carbonate crystals
2008). are embedded in the otoconial layer (OL), a
loose fibrous network at the top of the mem­
brane (Fig. 4.21A,B). In mammals the crystals
Mac ro m echan i cs and t he Otoco n i al
are in the form of calcite (Carlstrom 1963; Ross
M e m b ra n e
and Pote 1984) and are formed around organic
The otolithic or otoconial membrane (OM) is matrices. In each otolith organ, the crystals have
a flattened gelatinous structure, approximately a characteristic pattern (see Fig. 2. 7), which is
4. PHYS I OLOGY OF T H E VESTI B U LAR ORGANS 1 01

laid down during development. Because the lar meshwork (Dahlman 1971; Lim 1984). Given
crystals are denser than the surrounding endo­ its anisotropic structure, the CL is likely to be
lymph, the otoconial layer is displaced when the most compliant of the three layers in shear.
acted on by linear forces such as occur in a grav­ We consider the influence of otoconial mem­
itational field or when the head is accelerated in brane structure on response dyn amics and direc­
a straight line. Much remains to be learned about tional properties, beginning with the latter topic
the otoconia, including their initial formation, and concentrating on the insensitivity of the
the cues responsible for their elaborate pattern­ maculae to compressional forces. The OM con­
ing in each organ, and their maintenance during tains glycoproteins and proteoglycans ( Goodyear
life (Lundberg et al. 2006; Thalmann et al. 2001; and Richardson 2002) and can be described as a
Zhao et al. 2007). fibrous nervvork embedded in a hydrated gel. In
Lying underneath the OL is the gelatinous addition, hair bundles are sequestered in clear­
layer (CL), a relatively rigid structure consisting ings or tunnels in the CL (Fig. 4.21, middle)
of a dense, randomly oriented, cross-linked fila­ (Kachar et al. 1990; Lim 1984; but see Ross et al.
mentous network (Fig. 4.21D) that couples 1987). These structural features can explain the
motion of the OL to underlying structures. The insensitivity of the otolith organs to non-shearing
lowest or columnar layer (CL) (Fig. 4.2 1C) is a forces. Because hydrated gels have a turgor pres-
looser meshwork of vertically arranged filaments, . sure, they can resist compression (Alberts et al.
anchored to the apical surfaces of supporting 2002). At the same time, the fibrous network of
cells and sometimes referred to as the subcupu- the CL should attenuate the effects of traction.

Figure 4 .2 1 The otoconial membrane (OM) is composed of three layers.


Middle: Schematic of the three layers: OL, otoconial layer; CL, gelatinous
layer; CL, columnar layer. Shearing displacements of the CL are communi­
cated to the stereocilia (S) by way of the kinocilium (K) . A-D. Freeze-etch
specimens of the three layers. A, B. The OL consists of CaC03 crystals embed­
ded in a loose meshwork. Only the meshwork is seen here. C. In the CL, the
fibers are oriented so as to be compliant to shearing, but not to compressional,
forces. D. The CL consists of a dense meshwork of randomly oriented fibers
that give it rigidity in all directions. (From Kachar et al. 1990.)
1 02 TH E VESTI B U LAR SYSTEM

Sequestering the hair bundles in fluid compart­ lateral extent of the macula is so much larger
ments should contribute to their shielding from than the thickness of the OM that edge effects
compressional forces. can be ignored. The viscous drag exerted on the
Turning to response dynamics, both direct top surface of the O L by the endolymph is
observations of otoconial motion (de Vries 1950) neglected because it is much too small to result
and afferent recordings (Angelaki and Dickman in overdamped dyn amics.
2000; Fernandez and Goldberg 1976c) indicate To evaluate Equation 4. 10 requires a value for
that the mechanics are overdamped. To prevent 11p. We will use ti p = 0.33 gm/cm3 , a value mea­
underdamped oscillations requires much more sured by Steinhausen (1935) . A higher value of
damping than can be provided by the viscous tip was reported by Trincker (1962) . Setting
drag exerted by the endolymph ( Grant and Best a = 0.5 makes the CL incompressible. One
1986), suggesting that the OM itself provides the approach to estimating E is to consider steady­
necessary damping by being viscoelastic (i.e., state displacements by setting the time deriva­
combining the properties of a viscous fluid and tives in Equation 4. 10 to zero and Ax (t) to Ax ,
an elastic solid) . This is the expected mechanical a constant linear force. The result is
behavior of a hydrated gel. A one-dimensional
2
a2 X 211pAx (1 + O')
model incorporating this idea has been devel­ (4. 1 1)
oped (Grant and Cotton 1990; Grant et al. 1994; az E
Kondrachuk 2001b, 2002). The model includes \vith boundary conditions X = 0 at the neuroepi­
two simplifications: (1) the OM is considered as thelial surface (Z = O) and ax I az = 0 at the
lying in a single plane and (2) the CL and CL are top of the CL (Z = h). E can then be evaluated
considered as a single layer, which is referred to after a double integration of Equation 4. 1 1 to
as the CL. Once the basic equations are evalu­ yield
ated, we can consider the influence of the two
(l + <J)Ax • �p • h 2 [2(Z lh) - (Z lh) 2 ]
X(Z)
simplifications. (Kondrachuk 20016), E= (4. 12)
Our treatment follows that of Kondrachuk
(2001b) beginning with the second-order partial Studies of solitary hair cells suggest that maxi­
differential equation, mal transduction occurs when hair bundles of 10
2 E - a2 x
µm height are deflected in the excitatory direc­

ataz2 2(1 + O') az2 (4. 10)


l1p 2
a x --
J 3 x -- tion by 1 µm-that is, X(Z = 10 µm) = 1 µm
- + µ +
at (Holt et al. 1997). From afferent recordings
= 11p • Ax (t) (Fernandez and Goldberg 1976b), discharge
in the excitatory direcli_on typically saturates
X is the shearing displacement of the GL and Z for linear forces near A = 4 g (see Fig. 4.25).
its height measured from the surface of the neu­ Substituting these values into Equation 4. 12
roepithelium. Otoconial displacement is the result provides an E = 100 dyne/cm2 •
of a difference in density, tip = PoL - PENvo , Because Equation 4. 10 is second-order in
between the O L and the surrounding endo­ its time derivatives, it is governed by two time
lymph. Ax (t) = g x (t ) - ax (t) is referred to as the constants, r1 = 2µ(1 + 0") / E and r2 = tiph 2 I µ.
gravitoinertial force ( GIF) and is the sum of the An estimate of µ is provided by afferent-nerve
gravitational force [gx (t)] and the oppositely recordings, which suggest that the response
directed linear acceleration [-¾(t)] acting in the dynamics of the otolithic membrane are heavily
X or shearing direction. The left side of equation damped with a lower corner frequency near
4. 10 consists of an inertial term and two restor­ 5 Hz, corresponding to a time constant, 't { ':: 30
ing terms, one related to the GL's viscosity and ms. With E = 100 dyne/cm2 , this value of 't 1 gives
the other to its elasticity. The CL is characterized a value, µ = 1 poise, about 100 times the viscosity
by a coefficient of viscosity (µ), an elastic (Young's) of water (or endolymph). The system is heavily
modulus (E), and a Poisson's ratio (a). The model damped because 't2 ,:::: 3 x 10-3 ms, much smaller
is one-dimensional as the macula does not than 't 1 . 't2 is so small, in fact, that the OM
respond to compressional forces (Fernandez and should behave as a first-order system up to very
Goldberg 1976b; Kondrachuk 2001a) and the high frequencies (> 1,000 Hz) with response
4. PHYS I O LOGY OF THE VESTI B U LAR ORGANS 1 03

A B
0


0)


E 0.1 30
Q)
0)

c Q)

cu
Q) en
E
Q) if -60
cu
% 0. 0 1
0
-9 0

0.1 10 1 00 1 00 1 000
Frequency, Hz
0.01 1 000 0.01 0.1 10

Figure 4.22 Bode plot, viscoelastic model of otoconial membrane, including


gain (A) and phase (B) re gravitoinertial force. Lower comer frequency, 5 Hz.
Upper comer frequency, >>1000 Hz. Vertical dashed line divides the spec­
trum in low- and mid-frequency regions in which the output parallels sinusoidal
force (acceleration) and its integral (velocity), respectively.

being proportional to linear acceleration up to assume that virtually all of the shear deformation
5 Hz and to linear velocity at higher frequencies takes place in the CL. The main effect is to
(Fig. 4.22). increase the calculated values of E and µ and to
While the treatment is likely to be qualita­ reduce the overall displacement of the OM.
tively correct, its quantitative properties must be Another simplifying assumption is that each
viewed cautiously. For one thing, the estimate of macula lies in a single plane. This is clearly not
't 1 was obtained by assuming that the response the case as both organs show curvature. How
dynamics of regular afferents ·are determined this affects directional properties will depend on
solely by OM macromechanics. As afferent the mechanical properties of the GL. Were the
response dynamics may include phase leads GL perfectly rigid, the OM should move as a
introduced by transduction mechanisms subse­ unit and polarization vectors should all lie in a
quent to OM displacement, 't1 could be underes­ single plane. If, on the other hand, the CL were
timated. Furthermore, the response dynamics perfectly compliant, each point in the OM should
were observed only up to 2 Hz; a study of higher behave independently with its polarization vector
frequencies might be revealing. Even more being parallel to the local plane. The issue has
valuable would be direct observation of OM been explored with finite-element models
mechanics, which has been done only on (Jaeger et al. 2002; Kondrachuk 2001a), which
large saccular otolith of fish (de Vries 1950). suggest that vectors are predominantly influ­
Preliminary direct observations in mammals enced by local orientation. Obviously, these
suggests that the estimate of 't1 from afferent dis­ issues bear on the complementary roles played
charge is much too large (J.W. Grant, personal by the utricular and saccular maculae. In a previ­
communication); should this be confirmed, oto­ ous section, we suggested that the utricular
lith organs would remain acceleration sensors to macula encoded information about linear forces
very high frequencies. In the derivation, we have acting in the horizontal plane, whereas the sac­
assumed that stiffness resides in the OM and cular macula handled vertically oriented forces.
have ignored the possibly major contribution of If, in fact, directional properties are mainly influ­
hair bundles (Benser et al. 1993). enced by local orientation, units innervating the
We have also assumed that the GL and CL anterior, upwardly curving portion of the utricu­
had identical properties. Given the differences lar macula should be sensitive to vertical forces.
in their structure, it would seem reasonable to Given the curvature of the saccular macula
1 04 T H E VESTI B U LAR SYSTEM

(Lindeman 1969), some of its units might be from de to 2 Hz. Responses are nearly in phase
sensitive to horizontally oriented forces. In short, with linear acceleration, small phase leads being
while the major planes of the two maculae have replaced by larger phase lags as frequency is
complementary directional properties, both increased. The low-frequency phase lead may
organs could overlap in their handling of hori­ be due to adaptive processes, at least partly
zontal and vertical forces. located in the afferent terminal, while the high­
frequency phase lag is interpreted, ,vith some
reservations, as reflecting macromechanics. In
Afferent Response Dyn a m i cs
contrast to the relatively flat response dynamics
Because the utricular and saccular maculae of regular units, there is an approximately 10-fold
have similar structures, we might expect the two frequency-dependent increase in gain for irreg­
organs to have similar response dynamics. This ular units and conspicuous phase leads at all fre­
has been confirmed in afferent recordings quencies. These latter effects can be explained
(Fernandez and Goldberg 1976c). For either by the irregular units, besides sharing the adap­
organ, there are differences between regular tive and mechanical processes of regular units,
and irregular otolith afferents. As was the case being sensitive to the velocity of otolith displace­
for the cristae, regular otolith afferents have ment, as well as to the displacement itself. Were
response dyn amics resembling those expected of the response simply proportional to otolith veloc­
macromechanics, while irregular units show ity, we would expect that at some point the gain
large discrepancies from expectations. would increase linearly with stimulus frequency
Bode plots for the two kinds of otolith affer­ (f) and the phase would approach 90 degrees.
ents are shown in Figure 4.23. For regular units, Neither is the case. Rather, gain increases as the
the gain re linear acceleration is almost constant one-third power of frequency (£113 ) and phase

• --------

. //
... •
Regular I rregular
1 000 1 000
A 40 8 ''
40
'
./ / • 30
/ /

30
••
/
/
/

II!!/ �\\
\
'
/

20
/

20

• •
ci,
(/)

�Q)
10

'! ___ �----- , -■


ul
10
� • ____
- - - - - - - ,
Ol
U)
Q)


·15..
2
Q)
·ca
C -> ...._ _ '- - - - - - 1 00 0
1 00 - - - - - - - - - - - - - -
Cl)
0
''
'' • • co
11..
' ..c
CJ
••

I '\
-1 0
• • -1 0

•+- • •
\�
\

----
-20 -20

-30 •+- Gain


-30

■----■ Phase -40


rt.
-40

� ..... � .....
10 10 rf-
(0 L{) L{) L{) '"-: L{) C\J 0 (0 L{)
C\J
L{)
C\J
L{) '"-: L{) C\J
C\J C\J
0 0 q
Q
0 q 0
q
0 "-! q
q
0
q

Figure 4.2 3 Response dyn amics for two otolith afferents in the squirrel
monkey. Gains (left ordinate, solid line) and phases (right ordinate, dashed
line) re linear force versus sinusoidal frequency for a regular unit (A) and an
irregular unit (B). The irregular unit has more phasic response dynamics as
indicated by its larger phase leads and its larger high-frequency gain enhance­
ment. Curves are from empirical transfer functions, with arrows indicating
predicted static (DC) gains. Dashed horizontal lines indicate zero phase.
(From Fernandez and Goldberg 1976c. With permission of the American
Physiological Society.)
4. PHYSIOLOGY OF TH E VESTI B U LAR ORGANS 1 05

straddles 30 degrees. These are hallmarks of a larger during the faster (Fig. 4.24C) than during
fractional-differential operator ( Oldham and the slower force application (Fig. 4.24D) . Were
Spannier 2006; Thorson and Biederman-Thorson response proportional to velocity, we would
1974). expect an instantaneous step increase in dis­
Response dynamics can also be illustrated charge during the ramp and an instantaneous
by the responses to linear-force trapezoids. step fall on ramp termination. Both the increase
Trapezoids are presented that differ in the dura­ in discharge during the ramp and its decline
tions of their ascending and descending ramps, during the plateau are more gradual than this,
but not in their plateau duration or magnitude. features that are consistent vvith a fractional
For a regular unit (Fig. 4.24A,B), the peak operator.
responses are almost identical regardless of the
velocity of force application, illustrating that the
Dyn a m i c Ran ge of Affe re nt D i scharge
unit is almost exclusively sensitive to the instan­
taneous force. In contrast, an irregular unit shows To determine the dynamic ranges of otolith
velocity sensitivity: peak response is considerably afferents, responses to centrifugal forces over a

.,......,.-�
l
A B
.._.:,..-..,...:.,-:.11.·
���
-�
• - • • - ,I

·-z:::-..,.,,._�-
75 75
w -■�
.;.-■ �--

_.....
):-�
��

_;-----\_
50 50

Cl>
0 5 10 0 5 10 15

0 C D
1 00
1 00

75
75

50 50

25 25

0 5 10 0 5 10 15

Time, s

Figure 4.24 Responses t o linear force for a regular unit (A, B) and an irregular
unit ( C, D). Two force trapezoids whose ramps differ in their durations and
slopes are presented to each unit. The regular unit responds to force magni­
tude, but not to the rate of force application. Two components are seen in the
irregular unit, one proportional to the rate of force application, the other to
force magnitude. Curve in C is the response predicted from a transfer function
with a fractional operator. (From Fernandez and Goldberg 1976c. With per­
mission of the American Physiological Society. )
1 06 TH E VESTI B U LAR SYSTEM

......
A E
300
• ••
Q) •
'K 200
!!!.,
Q)

.c
(.)

t5 1 00 1 00 •


••• • • • • •
0 0

0 10 20 30 0 10 20 30 -5 0 5
Time (s) Force (xlg)

Figure 4.2 5 Responses of a regular otolith afferent to force trapezoids.


A. Excitatory response profiles: 1.23 g (_.); 2.46 g (O); and 4.92 g (e) force
plateaus. B-D. Inhibitory response profiles for 1.23, 2.46, and 4.92 g
force plateaus, respectively. E. Force-response relation. Note that the unit
continues to fire during inhibitory saturation. (From Fernandez and Goldberg
19766. With permission of the American Physiological Society.)

range of ±4.92 g were studied (Fernandez and to a dynamic range of 4.0 g. Average values
Goldberg 1976b). This was done only in regular (±SD) based on all the units in Figure 4.26 are
afferents. An example is seen in Figure 4.25. -0.58 ± 0.77 g (10%), 4.0 ± 0.96 g (90%), and 4.6
Responses are asymmetric. As excitatory force is ± 1. 1 g (dyn amic range). The dynamic range in
increased, responses continue to grow with dis­
charge approaching 300 spikes/s (Fig. 4.25A). In
contrast, even modest inhibitory responses satu­
rate so that increasing force does not reduce dis­
charge below a residual discharge of 17 spikes/s 300
(Fig. 4.25B-D) . The overall input-output rela­
tion is sigmoidal (Fig. 4.25E) . Maximum sensi­
·o..
Q)

tivity occurs at an inflection points displaced Cl)


2.4 g in the excitatory direction from zero force. 200
Cl)
C
The range of ±1 g is located in the lower, con­
cave upward part of the relation and is charac­ Q)

terized by a sensitivity of 28 spikes•s- 1/g, some 1 00


40% of the maximum sensitivity.
Data for this and other units are well fit by
first-order Boltzmann functions (Fig. 4.26)
1 -5 0 5
r(x) = ------ (4. 13)
1 + exp[-(x - x 0 ) / s] Force, xlg

where r(x) is the normalized response to a force Figure 4.2 6 Relations between response and force.
(x) , x0 is the force at the inflection point, and s is Each thin curve is the best-fitting Boltzmann relation
(Equation 4.13) for an individual, regularly discharging
unit. Responses are measured from the minimum rate
a sensitivity factor. The dynamic range can be
obtained during large inhibitory forces. Thick lines
defined by the two points along the stimulus axis
leading to 10% and 90% of the maximum indicate relation for unit illustrated in Figure 4.24
response. For the particular unit of Fig. 4.25, the (- - -) and relation based on average parameters for all
points are 0.34 g (10%) and 4.4 g (90%), leading units (-). (Based on Fernandez and Goldberg 19766.)
4. PHYSIOLOGY OF THE VESTI B U LAR ORGANS 1 07

the larger population extends well beyond ±1 g, are irregular, whereas most peripheral extrastri­
the range needed to encode static tilts. At the olar dimorphs are regular. Juxtastriolar dimorphs,
same time, there is considerable variation in the lying adjacent to the striola, are neither as regu­
maximal discharge of individual units. lar as the most regularly discharging units in the
peripheral extrastriola nor as irregularly dis­
charging as some striolar dimorphic or calyx
Variations i n G a i n and Phase
units. The relation behveen linear-force gain
As already noted, regular and irregular units (g2 H) and discharge regularity (cv* ) and that
differ in their response dynamics (see Figs. 4.23 between the phase of the linear-force response
and 4.24). This is reflected by the fact that ( <I> 2H) and cv* were similar for the extracellular
both the 2-Hz gain (g2H) and phase (<p2H) , (Fig. 4.27A,B) and intra-axonal samples (Fig.
obtained from extracellularly recorded units 4.27C,D). There is a single, semilogarithmic
(Fig. 4.27 A,B), increase with cv*. Although simi­ relation between phase and cv* for the intra-ax­
lar trends are seen in canal units (see Fig. 4. 15), onal sample, including dimorphic and calyx
there are obvious differences between the two units. In addition, a strong, nearly linear relation
data sets. As was true for canal units, there is a exists in the intra-axonal sample between g2 Hz
single, semilogarithmic relation between phase and cv* for regular dimorphic units (cv* � 0. 1).
and cv* (Fig. 4.27B). But unlike the case for Many irregular units have relatively high gains,
canal units, g2Hz for irregular utricular units does but none of the gains are as high as would be
not fall into t\vo discrete clusters (Fig. 4.27A). suggested from an extrapolation of the power­
Another difference may be noted. In both kinds law for regular units. In addition, there is no
of organs, the response dynamics of regular clear separation in the values of g2H2 for calyx and
and irregular afferents deviate from each other. irregular dimorphic units. This is in contrast to
The deviations fall into distinct low- and high­ data from the semicircular canals in which calyx
frequency ranges for canal units (see Fig. 4. 12 units have distinctively low 2-Hz gains. The low
and 4. 13) but are broadly distributed across the gains of calyx units in the canals suggested that
frequency range for otolith units (Fig. 4.23). synaptic inputs from type I hair cells were
smaller than expected from the number of syn­
aptic contacts with calyx inner faces. In an
Affe rent M o r p h o l ogy and Phys i o logy
attempt to retain the hyp othesis for the macula,
Intra-axonal labeling studies have been done it has been suggested that the comparable 2-Hz
for utricular afferents in the chinchilla ( Goldberg gains of utricular calyx and irregular dimorphs
et al. 1990b). For the labeled afferents, polariza­ can be explained by the former units having
tion vectors were determined by static tilts. The a somewhat more irregular discharge and
locations and vectors of the labeled units were more phasic response dynamics (for details, see
compared with published morphological-polar­ Goldberg et al. 1990b).
ization maps (see Fig. 2. 10). The great majority \Ve can summarize the results as follows.
of labeled units had vectors consistent with the Calyx afferents, which are confined to the striola,
maps. As expected from extracellular labeling are the most irregularly discharging afferents,
studies (see Fig. 2. 15), in!ra-axonally labeled and their responses to 2-Hz sinusoidal linear
calyx units were found exclusively in the striola, forces are characterized by high gains and the
while dimorphic units were found in the striola, largest phase leads. Dimorphic afferents, which
juxtastriola, and extrastriola. None of the labeled are found throughout the macula, provide
units were of the bouton variety. Their absence evidence for a regional variation in discharge
from the sample is hardly surprising considering properties. Striolar dimorphs are distinguished
that they have small axons and make up only from extrastriolar dimorphs in having a more
z 10% of the total innervation. irregular discharge, higher gains, and larger
Calyx units are the most irregularly discharg­ phases. In all these respects, juxtastriolar
ing units in the sample (Fig. 4.27C,D). The dis­ dimorphs show intermediate properties. As
charge regularity of dimorphic units depends on already noted, no physiologically characterized
their location. Most of those located in the striola bouton fibers were recovered. Based on their
........ � .... .....
1 08 THE VESTI B U LAR SYSTEM

'. : .. . . �.. :. .
1 000 C

. .'
A

9l
en
Q)

·o.. 1 00
• • • •
..
I•


·ca
.!E,,

...
C

(9

..
0

10


80 80
B D
60 60

40 40

Q)
0)

en
Q) 20 20
a..
..c
0 0

-20 -20 Calyx •


Dimorphic o
U n labled
-40 -40 -1----------.----.-...........................
.0 1
Coefficient of variation
.01

Figure 4.2 7 Gains and phases versus normalized coefficient of variation (cv* ).
Utricular afferents in the chinchilla responding to 2-Hz sinusoidal centrifugal
forces parallel to each unit's polarization vector. Each point represents one
unit. A, B. Extracellularly recorded units. Sloping line in A is best-fitting power
law relation betvv'een gain and cv* for units, cv* < 0.18. Horizontal line indi­
cates average gain for units, cv* > 0.18. Straight line in B is best-fitting semi­
logarithmic relation behveen phase and cv* for all units. C, D. Corresponding
data for intracellularly labeled units, including 24 dimorphic and 9 calyx units
(see Key). Unlabeled units are shown as small dots. Calyx units are the most
irregularly discharging afferents and have the largest phase leads; their gains
are not distinctive. Gains and phases of dimorphic units increase in association
with cv*; regular units are found in the extrastriola and irregular units in the
striola. (From Goldberg et al. 1990b. With permission of the American
Physiological Society.)

location in the peripheral extrastriola, they might from the number of synapses they make with
be expected to have a regular discharge, low hair cells.
gains, and small phases. In the canals, calyx We now consider a functional correlate of the
afferents have distinctively low gains compared difference between the canal and otolith calyx
to irregular dimorphs also located in the afferents. In the case of the canals, it has been
central zone. This is not the case for utricular suggested that the gains of calyx afferents would
macula, but the results do not disprove the prop­ allow these irregular units to respond linearly to
osition that calyx endings are associated with a even very rapid he�d saccades. In contrast, the
lower synaptic gain than would be predicted linear accelerations occurring during everyday
4. PHYSIOLOGY OF TH E VESTI B U LAR ORGA NS 1 09

activity are of limited magnitude, less than 1 g for elasticity. Throughout a broad frequency range
walking and less than 2 g for running and hop­ (0.025 to 25 Hz), viscosity dominates and cupu­
ping (Pozzo et al. 1990). Similar stimulus magni­ lar displacement is proportional to angular head
tudes occur during non-locomotor activities velocity. The response properties of dimorphic
(MacDougall and Moore 2005). Maximal gains of units vary with location. Those in the peripheral
irregular otolith afferents, whether calyx or dimor­ zone are regularly discharging and have low rota­
phic, are near 300 spikes•s- 1/g- 1 for high-frequency tional gains and tonic response dynamics. Bouton
linear forces (Angelaki and Dickman 2000; units are located in the peripheral zone of the
Fernandez and Goldberg 1976c; Goldberg et al. crista and may be presumed to resemble their
1990a) (see Fig. 4.23). Peak excitatory responses dimorphic neighbors. Calyx units, which inner­
should be on the order of 400 spikes/s, within the vate the central zone of the crista, are the most
upper response limits of many otolith afferents. irregular and most phasic afferents and have
lower gains for mid-frequency rotations than do
similarly located irregular dimorphs.
4.4 SUMMARY The directional properties of each otolith
afferent can be summarized by a functional
Semicircular canals provide a three­ polarization vector. Since otolith organs respond
dimensional reconstruction of angular forces to shearing but not to compressional forces, most
acting on the head, while otolith afferents do the utricular afferents respond to linear forces in
same for linear forces. Afferents have a resting the horizontal plane, whereas most saccular
discharge, whose presence allows bidirectional afferents are sensitive to dorsoventral forces.
responses, effectively eliminates the existence of The biomechanics of the otoconial membrane
a sensory threshold, and provides a tonic excit­ acts like a first-order system that is sensitive to
atory input to central vestibular pathways. Fibers, linear acceleration below a corner frequency
whether they innervate canal or otolith organs, that still needs to be determined and to linear
differ in discharge regularity, which is largely velocities above it. Striolar otolith afferents are
determined by the physiology of the spike irregularly discharging and phasic and have high
encoder located in the postsynaptic terminal. gains to linear 'accelerations, whereas peripheral
Irregular afferents are more sensitive than are extrastriolar units are regularly discharging and
regular afferents to afferent and efferent synap­ tonic and have low gains. Striolar calyx units are
tic inputs and to externally applied galvanic stim­ more irregular and more phasic than similarly
ulation. Response dynamics of regular afferents located dimorphic afferents. Unlike calyx units
resemble the expected biomechanics of cupular in the semicircular canals, those in the utricular
or otolithic displacement, whereas irregular macula do not have distinctively low gains. From
afferents show more phasic response dynamics. a functional perspective, the difference in calyx
Differences in response dynamics, while corre­ gains may be related to the range of stimulus
lated with discharge regularity, cannot be magnitudes handled by canal and otolith
explained by variations in encoder sensitivity. organs.
Each semicircular canal responds to head
rotations having a component near the canal
plane. The discharge of all afferents innervating 4.5 SELECTED READ I NGS

Eatock RA, Lysakowski A (2006) Mammalian vestibu­


a crista is increased by rotations in one and the
lar hair cells. In: Vertebrate Hair Cells (Eatock RA,
same direction and decreased by rotations in the
opposite direction. Coplanar canal pairs on the Fay RR, Popper AN, eds), pp. 348-442. Berlin:
two sides have opposite directional properties. It Springer-Verlag.
Goldberg JM (2000) Afferent diversity and the orga­
nization of central vestibular pathways. Exp Brain
is the difference in activity of coplanar canal pairs
Res 130:277-297.
that is interpreted by the brain as a rotation. The
Highstein SM, Rabbitt RD, Holstein GR, Boyle RD
biomechanics of the semicircular canals follows
that of an overdamped torsion pendulum whose (2005) Determinants of spatial and temporal coding
response dynamics incorporates the effects of by semicircular canal afferents. J Neurophysiol
endolymph inertia and viscosity and of cupular 93:2359-2370.
1 10 THE VESTI B U LAR SYSTEM

Lewis ER, Leverenz EL, Bialek WS (1985) The Benser ME, Issa NP, Hudspeth AJ (1993)
Vertebrate Inner Ear. Boca Raton FL: CRC Hair-bundle stiffness dominates the elastic reac­
Press. tance to otolithic-membrane shear. Hear Res
Lysakowski A, Goldberg JM (2004) Morphophysiology 68:243-252.
of the vestibular periphery. In: The Vestibular Blanks RH, Curthoys IS, Bennett ML, Markham CH
System (Highstein SM, Popper A, Fay RR, eds), (1985) Planar relationships of the semicircular
pp. 57- 152. New York: Springer-Verlag. canals in rhesus and squirrel monkeys. Brain Res
Rabbitt RD, Damiano ER, Grant JW (2004a) 340:315-324.
Biomechanics of the semicircular canals and oto­ Blanks RH, Curthoys IS, Markham CH (1972) Planar
lith organs. In: The Vestibular System (Highstein relationships of semicircular canals in the cat. Am J
SM, Popper A, Fay RR, eds), pp. 153-201. Physiol 223:55-62.
New York: Springer-Verlag. Blanks RH, Precht W (1978) Response properties of
vestibular afferents in unanesthetized cats during
optokinetic and vestibular stimulation. Neurosci
Lett 10:225-229.
REFERENCES Borst A, Theunissen FE (1999) Information theory
and neural coding. Nat Neurosci 2:947-957.
Alberts BM, Roberts K, Lewis J, Raff M, Walter P, Boyle R, Carey JP, Highstein SM (1991) Morphological
Johnson A (2002) Molecular Biology of the Cell, correlates of response dynamics and efferent stim­
4th ed. New York: Garland Science. ulation in horizontal semicircular canal afferents of
Anastasio TJ, Correia MJ, Perachio AA (1985) the toadfish, Opsanus tau. J Neurophysiol 66:1504-
Spontaneous and driven responses of semicircular 1521.
canal primary afferents in the unanesthetized Boyle R, Highstein SM (1990) Resting discharge and
pigeon. J Neurophysiol 54:335-347. response dynamics of horizontal semicircular canal
Anderson JH, Blanks RHI, Precht W (1978) Response afferents in the toadfish, Opsanus tau. J Neurosci
characteristics of semicircular canal and otolith 10:1557-1569.
systems in the cat. I. Dynamic responses of Bronte-Stewart HM, Lisberger SC (1994)
primary vestibular fibers. Exp Brain Res 32: Physiological properties of vestibular primary
491-507. afferents that mediate motor learning and normal
Angelaki DE (1992) Two-dimensional coding of linear performance of the vestibulo-ocular reflex in mon­
acceleration and the angular velocity sensitivity of keys. J Neurosci 14: 1290-1308.
the otolith system. Biol Cybern 67:511-521. Breuer J (1874) Ueber die Function der Bogengange
Angelaki DE, Bush GA, Perachio AA (1993) Two­ des Ohrlabyrinths. Wien Med Jahrb: 72-124.
dimensional spatiotemporal coding of linear accel­ Breuer J (1875) Beitrage zur Lehre von statischen
eration in vestibular ·nuclei neurons. J Neurosci Sinne. Wien Med Jahrb:87-156.
13:1403-1417. Brichta AM, Goldberg JM (2000a) Morphological
Angelaki DE, Dickman JD (2000) Spatiotemporal identification of physiologically characterized affer­
processing of linear acceleration: primary affer­ ents innervating the turtle posterior crista. J
ent and central vestibular neuron responses. Neurophysiol 83:1202-1223.
J Neurophysiol 84:2113-2132. Bush GA, Perachio AA, Angelaki DE (1993) Encoding
Angelaki DE, Shaikh AG, Green AM, Dickman JD of head acceleration in vestibular neurons. I.
(2004) Neurons compute- internal models of the Spatiotemporal response properties to linear accel­
physical laws of motion. Nature 430:560-564. eration. J Neurophysiol 69:2039-2055.
Armand M, Minor LB (2001) Relationship between Calabrese DR, Hullar TE (2006) Planar relationships
time- and frequency-domain analyses of angular of the semicircular canals in tv,o strains of mice. J
head movements in the squirrel monkey. J Comput Assoc Res Otolaryngol 7:151-159.
Neurosci 11:217-239. Carlstrom D (1963) A crystallographic study of verte­
Baird RA (1994) Comparative transduction mecha­ brate otoliths. Biol Bull 125:441-463.
nisms of hair cells in the bullfrog utriculus. II. Clark B (1967) Thresholds for the perception of
Sensitivity and response dynamics to hair bundle angular acceleration in man. Aerospace Med
displacement. J Neurophysiol 71:685-705. 38:443-450.
Baird RA, Desmadryl G, Fernandez C, Goldberg JM Coats AC, Smith SY (1967) Body position and the
(1988) The vestibular nerve of the chinchilla. II. intensity of caloric nystagmus. Acta Otolaryngol
Relation between afferent response properties and (Stockh) 63:515-532.
peripheral innervation patterns in the semicircular Correia MJ, Perachio AA, Dickman JD, Kozlovskaya
canals. J Neurophysiol 60:182-203. IB, Sirota MG, Yakushin SB, Beloozerova IN
Barany R (1906) Untersuchungen uber den vom­ (1992) Changes in monkey horizontal semicircular
Vestibularapparat des Ohres reflektorisch aus­ canal afferent responses after spaceflight. J Appl
gelosten rythmischen Nystagmus und seine Physiol 73:112S-120S.
Begleiterscheinungen. Mschr Ohrenheilkd 40: Crum Brown A (1874) On the semicircular canals of
193-297. the internal ear. J Anat Physiol.
4. PHYSIOLOGY OF TH E V ESTI B U LAR ORGANS 111

Curthoys IS (1982) The response of primal) horizon­ Fernandez C, Goldberg JM, Abend WK (1972)
tal semicircular canal neurons in the rat and Response to static tilts of peripheral neurons inner­
guinea pig to angular acceleration. Exp Brain Res vating otolith organs of the squirrel monkey.
4 7:286-294. J Neurophysiol 35:978-987.
Curthoys IS, Curthoys EJ, Blanks RH, Markham CH Furukawa T, et al. (1982) Quantal analysis of a decre­
(1975) The orientation ofthe semicircular canals in mental response at hair cell -afferent fibre synapse
the guinea pig. Acta Otolaf)rngol 80:197-205. in the goldfish sacculus. J Physiol (Land) 322:
Curthoys IS, Oman CM (1986) Dimensions of the 181-195.
horizontal semicircular duct, ampulla and utricle in Furukawa T, Matsuura S (1978) Adaptive rundown of
rat and guinea pig. Acta Otolaf)'l1.gol (Stockh) excitatory post-synaptic potentials at synapses
101:1-10. between hair cells and eight nerve fibres in the
de Vries H (1950) The mechanics of the labyrinthine goldfish. J Physiol (Land) 276:193-209.
otoliths. Acta Otolaf)'l1.gol (Stockh) 38:262-273. Goldberg JM (2000) Afferent diversity and the orga­
Dingledine R, Borges K, Bowie D, Traynelis SF nization of central vestibular pathways. Exp Brain
(1999) The glutamate receptor ion channels. Res 130:277-297.
Pharmacol Rev 51:8-61. Goldberg JM, Brichta AM (20026) Functional
Dahlman CF (1935) Some practical and theoretical analysis of whole cell currents from hair cells
points in labyrinthology. Proc Roy Soc Med: of the turtle posterior crista. J Neurophysiol
1371-1380. 88:3279-3292.
Dahlman CF (1971) The attachment of the cupulae, Goldberg JM, Desmadryl G, Baird RA, Fernandez C
otolith and tectorial membranes to the sensory cell (1990a) The vestibular nerve of the chinchilla.
areas. Acta Otolaf)'l1.gol (Stockh) 71:89-105. IV. Discharge properties of utricular afferents.
Dunlap M, Spoon C, Grant JW (2011) Dynamic J Neurophysiol 63:781-790.
response of the otoconial membrane of the turtle Goldberg JM, Desmadryl G, Baird RA, Fernandez C
utricle. Assoc. Res. Otolaf)1l1gol. Abs. (in press). (19906) The vestibular nerve of the chinchilla.
Eatock RA, Lysakowski A (2006) Mammalian vestibu­ V. Relation between afferent discharge properties
lar hair cells. In: Vertebrate Hair Cells (Eatock RA, and peripheral innervation patterns in the utricular
Fay RR, Popper AN, eds), pp. 348--442. Berlin: macula. J Neurophysiol 63:791-804.
Springer-Verlag. Goldberg JM, Fernandez C (1971a) Physiology
Edwards CH, Penney DE (2000) Elementary of peripheral neurons innervating semicircular
Differential Equations with Boundary Value canals of the squirrel monkey. I. Resting discharge
Problems. Upper Saddle River, NJ: Prentice Hall. and response to constant angular accelerations.
Estes MS, Blanks RH, Markham CH (1975) J Neurophysiol 34:635-660.
Physiologic characteristics of vestibular first-order Goldberg JM, Fernandez C (19716) Physiology of
canal neurons in the cat. I. Response plane deter­ peripheral neurons innervating semicircular canals
mination and resting discharge characteristics. of the squirrel monkey. III. Variations among
J Neurophysiol 38:1232-1249. units in their discharge properties. J Neurophysiol
Ezure K, Cohen MS, Wilson VJ (1983) Response of 34:676-684.
cat semicircular canal afferents to sinusoidal polar­ Goldberg JM, Fernandez C (1975) Responses
izing currents: implications for input-output prop­ of peripheral vestibular neurons to angular and
erties of second-order neurons. J Neurophysiol linear accelerations in the squirrel monkey. Acta
49:639-648. Otolaf)'l1.gol (Stockh) 80:101-110.
Fernandez C, Goldberg JM (1971) Physiology of Goldberg JM, Fernandez C (1977) Conduction times
peripheral neurons innervating semicircular canals and background discharge of vestibular afferents.
of the squirrel monkey. II. Response to sinusoidal Brain Res 122:545-550.
stimulation and dyn amics of peripheral vestibular Goldberg JM, Fernandez C, Smith CE (1982)
system. J Neurophysiol 34:661-675. Responses of vestibular-nerve afferents in the
Fernandez C, Goldberg JM (1976a) Physiology of squirrel monkey to externally applied galvanic cur­
peripheral neurons innervating otolith organs of rents. Brain Res 252:156-160.
the squirrel monkey. I. Response to static tilts and Goldberg JM, Smith CE, Fernandez C (1984) Relation
to long-duration centrifugal force. J Neurophysiol between discharge regularity and responses to exter­
39:970-984. nally applied galvanic currents in vestibular nerve
Fernandez C, Goldberg JM (19766) Physiology of . afferents of the squirrel monkey. J Neurophysiol
peripheral neurons innervating otolith organs of 51:1236-1256.
the squirrel monkey. II. Directional selectivity Goltz F (1870) Ueber die physiologische Bedeutung
and force-response relations. J Neurophysiol der Bogengange der Ohrlabytinths. Pflugers Archiv
39:985-995. Eur J Physiol 3:172-192.
Fernandez C, Goldberg JM (1976c) Physiology of Goodyear RJ, Richardson GP (2002) Extracellular
peripheral neurons innervating otolith organs of matrices associated with the apical surfaces of
the squirrel monkey. III. Response dyn amics. sensory epithelia in the inner ear: molecular and
J Neurophysiol 39:996-1008. structural diversity. J Neurobiol 53:212-227.
1 12 THE VESTI B U LAR SYSTEM

Goutman JD, Glowatzki E (2007) Time course and Holt JC, Xue J-T, Brichta AM, Goldberg JM (20066)
calcium dependence of transmitter release at a Transmission between type II hair cells and bouton
single ribbon synapse. Proc Natl Acad Sci U S A afferents in the turtle posterior crista. J
104: 16341-16346. Neurophysiol 95:428-452.
Grant JW, Best WA (1986) Mechanics of the otolith Honrubia V, Hoffman LF, Sitko S, Schwartz IR (1989)
organ-dynamic response. Ann Biomed Eng Anatomic and physiological correlates in bullfrog
14:241-256. vestibular nerve. J Neurophysiol 61:688-701.
. Grant JW, Cotton JR (1990) A model for otolith Hullar TE, Della Santina CC, Hirvonen T, Lasker
dynamic response ·with a viscoelastic gel layer. J DM; Carey JP, Minor LB (2005) Responses of
Vestib Res 1:139-151. irregularly discharging chinchilla semicircular
Grant JW, Huang CC, Cotton JR (1994) Theoretical canal vestibular-nerve afferents during high­
mechanical frequency response of the otolithic frequency head rotations. J Neurophysiol 93:
organs. J Vestib Res 4:137-151. 2777- 2786.
Grossman GE, Leigh RJ, Abel LA, Lanska DJ, Hullar TE, Minor LB (1999) High-frequency dynam­
Thurston SE (1988) Frequency and velocity of ics of regularly discharging canal afferents provide
rotational head perturbations during locomotion. a linear signal for angular vestibuloocular reflexes.
Exp Brain Res 70:470-476. J Neurophysiol 82:2000-2005.
Gu�dry FE, Jr. (1974) Psychophysics of vestibular Hunter-Duvar IM, Hinajosa R (1984) Vestibule:
sensation. In: Handbook of Sensory Physiology. sensory epithelia. In: Ultrastructural Atlas of the
Volume VI. Vestibular System. Part 2: Psycho­ Inner Ear (Friedmann I, Ballantyne J, eds),
physics, Applied Aspects and General Inter­ pp. 211-244. London: Butterworths.
pretations (Kornhuber HH, ed), pp. 3-154. Berlin: Huterer M, Cullen KE (2002) Vestibuloocular
Springer-Verlag. reflex dynamics during high-frequency and high­
Guth PS, Perrin P, Norris CH, Valli P (1998) The acceleration rotations of the head on body in rhesus
vestibular hair cells: post-transductional signal monkey. J Neurophysiol 88:13-28.
processing. Prag Neurobiol 54:193-247. Igarashi M (1967) Dimensional study of the vestibular
Haque A, Angelaki DE, Dickman JD (2004) Spatial apparatus. Laryngoscope 77:1806-1817.
tuning and dynamics of vestibular semicircular Jaeger R, Takagi A, Haslwanter T (2002) Modeling
canal afferents in rhesus monkeys. Exp Brain Res the relation between head orientations and otolith
155:81-90. responses in humans. Hear Res 173:29-42.
Highstein SM, Rabbitt RD, Boyle R (1996) Jones GM, Spells KE (1963) A theoretical and com­
Determinants of semicircular canal afferent parative study of the functional dependence of the
response dynamics in the toadfish, Opsanus tau. semicircular canal upon its physical dimensions.
J Neurophysiol 75:575-596. Proc Roy Soc London B 157:403-419.
Highstein SM, Rabbitt RD, Holstein GR, Boyle RD Kachar B, Parakkal M, Fex J (1990) Structural basis
(2005) Determinants of spatial and temporal coding for mechanical transduction in the frog vestibular
by semicircular canal afferents. J N europhysiol sensory apparatus: I. The otolithic membrane.
93:2359-2370. Hear Res 45:179-190.
Hillman DE, McLaren JW (1979) Displacement con­ Kachar B, Parakkal M, Kure M, Zhao Y, Gillespie PG
figuration of semicircular canal cupulae. (2000) High-resolution structure of hair-cell tip
Neuroscience 4:1989-2000. links. Proc Natl Acad Sci U S A 97:13336-13341.
Hirsch-Shell D. Paulin M, Hoffman L (2011) Keller EL (1976) Behavior of horizontal semicircu­
The contributions of signal and noise to infor­ lar canal afferents in alert monkey during vestibu­
mation transmission rates in mammalian vesti­ lar and optokinetic stimulation. Exp Brain Res
bular afferent neurons. Assoc Res Otolaryngol 24:459-4 71.
Abs: 761. Kondrachuk AV (2001a) Finite element modeling of
Holstein GR, Martinelli GP, Henderson SC, Friedrich the 3D otolith structure. J Vestib Res 11:13-32.
VL, Jr., Rabbitt RD, Highstein SM (2004a) Kondrachuk AV (20016) Models of the dynamics of
Gamma-aminobutyric acid is present in a spatially otolithic membrane and hair cell bundle mechan­
discrete subpopulation of hair cells in the crista ics. J Vestib Res 11:33-42.
ampullaris of the toadfish Opsanus tau. J Comp Kondrachuk AV (2002) Otoliths as biomechanical
Neurol 47:1-10. gravisensors. Adv Space Res 30:745-750.
Holstein GR, Rabbitt RD, Martinelli GP, Friedrich Lewis ER, Leverenz EL, Bialek WS (1985) The
VLJ, Boyle R, Highstein SM (20046) Covergence Vertebrate Inner Ear. Boca Raton FL: CRC
of excitatory and inhibitory hair cell transmitter Press.
shapes vestibular afferent responses. Proc Natl Liao K, Kumar AN, Han YH, Grammer VA, Gedeon
Acad Sci U S A 101. BT, Leigh RJ (2005) Comparison of velocity wave­
Holt JR, Corey DP, Eatock RA ( 1997) Mechanoelectric forms of eye and head saccades. Ann N Y Acad Sci
transduction and adaptation in hair cells of the 1039:477-479.
mouse utricle, a low-frequency vestibular organ. J Lim DJ (1976) Morphological and physiological cor­
Neurosci 17:8739-8748. relates in cochlear and vestibular sensory epithelia.
4. PHYS I OLOGY O F T H E VESTI B U LAR ORGANS 113

In: Scanning Electron Microscopy/1976/11. Mach E (1874) Phvsicalische Versuche iiber den
Chicago: Illinois Institute of Technology Research Gleichgewichtssin'n des Menschen. Wien Sitzb
Institute. Kais Akad Wiss 69:121-135.
Lim DJ (1984) The development and structure of the Matano S (1986) A volumetric comparison of the ves­
otoconia. In: Ultrastructural Atlas of the Inner Ear tibular nuclei in primates. Folia Primatol (Basel)
(Friedmann I, Ballantyne J, eds), pp. 245-269. 47:189-203.
London: Butterworths. Melvill Jones G (1974) The functional significance
Lindeman HH (1969) Studies on the morphology of semicircular canal size. In: Handbook of
of the sensory regions of the vestibular appara­ Sensory Physiology. Volume VI. Vestibular System.
tus. Ergebnisse Anatomie Entwicklungsgeschite Part 1: Basic Mechanisms (Kornhuber HH, ed) ,
42:1-11.3. pp. 171-184. Berlin: Springer-Verlag.
Lisberger SC, Pavelko TA (1986) Vestibular signals Minor LB, Goldberg JM (1990) Influence of static
carried by pathways subserving plasticity of the head position on the horizontal nystagmus evoked
vestibulo-ocular reflex in monkeys. J Neurosci by caloric, rotational and optokinetic stimulation in
6:346-354. the squirrel monkey. Exp Brain Res 82:1-13.
Loe PR, Tomko DL, Werner G (1973) The neural Minor LB, Lasker DM, Backous DD, Hullar TE
signal of angular head position in primary afferent (1999) Horizontal vestibuloocular reflex evoked
vestibular nerve axons. J Physiol (Lond) 230: by high-acceleration rotations in the squirrel
29-50. monkey. I. Normal responses. J Neurophysiol
Louie AW, Kimm J (1976) The response of 8th nerve 82:1254-1270.
fibers to horizontal sinusoidal oscillation in the Myers SF, Lewis ER (1990) Hair cell tufts and affer­
alert monkey. Exp Brain Res 24:447-457. ent innervation of the bullfrog crista ampullaris.
Lowenstein 0, Sand A (1936) The activity of the hori­ Brain Res 534:15-24.
zontal semicircular canal of the dogfish, Scyllium Oldham KB, Spannier J (2006) The Fractional
canalicula. J Exp Biol 13:416-428. Calculus: Theory and Application of Differentiation
Lowenstein 0, Sand A (1940a) The individual and and Integration to Arbitrary Order Mineola NY:
integrated activity of the semicircular canals of the Dover.
elasmobranch labyrinth. J Physiol (Lond) 99: Olson ES (2001) Intracochlear pressure measure­
89-101. ments related to cochlear tuning. J Acoust Soc Am
Lowenstein 0, Wersa.11 J (1954) A functional interpre­ 110:349-367.
tation of the electron microscopic structure of the Oman CM, Marcus EN, Curthoys IS (1987) The
sensory hairs in the cristae of the elasmobranch, influence of semicircular canal morphology on
Raja clavata. Nature 184:1807-1810. endolymph flow dynamics. An anatomically
Lundberg YW, Zhao X, Yamoah EN (2006) Assembly descriptive mathematical model. Acta Otolaryngol
of the otoconia complex to the macular sensory ( Stockh) 103:1-13.
epithelium of the vestibule. Brain Res 1091: Oman CM; Young LR (1972) The physiological range
47-57. of pressure difference and cupula deflections in
Lysakowski A (1996) Syn aptic organization of the the human semicircular canal. Acta Otolaryngol
crista ampullaris in vertebrates. Ann N Y Acad Sci (Stockh) 74:324-331.
781:164-182. Paige GD (1985) Caloric responses after horizontal
Lysakowski A, Goldberg JM (1997) A regional ultra­ canal inactivation. Acta Otolaryngol (Stockh)
structural analysis of the cellular and synaptic 100:321-327.
architecture in the chinchilla cristae ampullares. J Perachio AA, Correia MJ ( 1983) Responses of
Comp Neurol 389:419-443. semicircular canal and otolith afferents to small
Lysakowski A, Goldberg JM (2004) Morphophysiology angle static head tilts in the gerbil. Brain Res
of the vestibular periphery. In: The vestibular 280:287-298.
system (Highstein SM, Popper A, Fay RR, eds), Plotnik M, Goldberg JM, Marlinski VV (1999)
pp. 57-152. New York: Springer-Verlag. Excitatory response-intensity relations in afferents
Lysakowski A, Goldberg JM (2008) Ultrastructural from the chinchilla's superior and horizontal canals.
analysis of the cristae ampullares in the squirrel Soc Neurosci Abstr 25:663.
monkey ( Saimiri sciureus). J Comp Neural 511: Plotnik M, Marlinski V, Goldberg JM (2005) Efferent­
47-64. mediated fluctuations in vestibular nerve discharge:
Lysakowski A, Minor LB, Fernandez C, Goldberg JM a novel, positive-feedback mechanism of efferent
(1995) Physiological identification of morphologi­ control. J Assoc Res Otolaryngol 6:311-323.
cally distinct afferent classes innervating the cristae Pozzo T, Berthoz A, Lefort L (1990) Head stabiliza­
ampullares of the squirrel monkey. J Neurophysiol tion during various locomotor tasks in humans.
73:1270-1281. I. Normal subjects. Exp Brain Res 82:97-106.
MacDougall HG, Moore ST (2005) Marching to the Rabbitt RD (1999) Directional coding of three­
beat of the same drummer: the spontaneous dimensional movements by the vestibular semicir­
tempo of human locomotion. J Appl Physiol cular canals.[erratum appears in Biol Cybern 2000
99:1164-1173. Apr;82( 4):355]. Biol Cybern 80:417-431.
I'

1 14 THE VESTI B U LAR SYSTEM

Rabbitt RD, Breneman KD, King C, Yamauchi AM, Smith CE, Goldberg JM (1986) A stochastic afterhy­
Boyle R, Highstein SM (2009) Dynamic displace­ perpolarization model of repetitive activity in
ment of normal and detached semicircular canal vestibular afferents. Biol Cybern 54:41-51.
cupula. J Assoc Res Otolaryngol 10:497-509. Somps CJ, Schor RH, Tomko DL (1994) Vestibular
Rabbitt RD, Damiano ER, Grant JW (2004a) afferent responses to linear accelerations in the
Biomechanics of the semicircular canals and alert squirrel monkey. In: NASA Technical
otolith organs. In: The Vestibular System (Highstein Memorandum 4581. Moffett Field, CA: Ames
SM, Popper A, Fay RR, eds), pp. 153-201. Research Center.
New York: Springer-Verlag. Spassova MA, Avissar M, Furman AC, Crumling MA,
Ramachandran R, Lisberger SC (2005) Normal Saunders JC, Parsons TD (2004) Evidence that
performance and expression of learning in the rapid vesicle replenishment of the synaptic ribbon
vestibulo-ocular reflex (VOR) at high frequencies. mediates recovery from short-term adaptation at
J Neurophysiol 93:2028-2038. the hair cell afferent synapse. J Assoc Res
Ramachandran R, Lisberger SC (2006) Transformation Otolaryngol 5:376-390.
of vestibular signals into motor commands in the Spoor F (1998) Comparative review of the human
vestibuloocular reflex pathways of monkeys. bony labyrinth. Yearbook of Physical Anthropology
J Neurophysiol 96:1061-1074. 41:211-251.
R3;_mprashad F, Landolt JP, Money KE, Laufer J Stahl JS (1992) Signal Processing in the Vestibulo­
(1984) Dimensional analysis and dynamic response ocular Reflex of the Rabbit (PhD thesis). New
characterization of mammalian peripheral vestibular York: New York University.
structures. Am J Anat 169:295-313. Steinhausen W (1931) Ueber den Nachweis der
Reisine H, Simpson JI, Henn V (1988) A geometric Bewegung der Cupula in der intakten
analysis of semicircular canals and induced activity Bogengansampulle des Labyrinthes bei der nati.ir­
in their peripheral afferents in the rhesus monkey. lichen rotatorischen und calorischen Reizung.
Ann N Y Acad Sci 545:10-20. Pflugers Arch 228:322-328.
Rieke F, Warland D, de Ruyter van Steveninck R, Steinhausen W (1935) Uber die durch die Otolithen
Bialek WS (1997) Spikes. Exploring the Neural ausgelosten Krafte. Pflugers Arch 235:538-544.
Code. Cambridge, MA: MIT Press. Stephan H, Frahm H, Baron G (1981) New and
Robles L, Ruggero MA (2001) Mechanics of the mam­ revised data on volumes of brain structures inin­
malian cochlea. Physiol Rev 81:1305-1352. sectivores and primates. Folia Primatol (Basel)
Ross MD, Komorowski TE, Donovan KM, Pote KG 35:1-29.
(1987) The suprastructure of the saccular macula. Taglietti V, Rossi ML, Casella C (1977) Adaptive dis­
Acta Otolaryngol (Stockh) 103:56-63. tortions in the generator potential of semicircular
Ross MD, Pote KG (1984) Some properties of otoco­ canal sensory afferents. 123: 41-57.
nia. Philosophical Trans Royal Soc London B Biol Thalmann R, Ignatova E, Kachar B, Ornitz DM,
Sci 304:445-452. Thalmann I (2001) Development and maintenance
Rushton WAH (1951) A theory of the effects of of otoconia: biochemical considerations. Ann N Y
fibre size in medullated ne�e. J Physiol (Lond) Acad Sci 942:162-178.
115:101-122. Thorson J, Biederman-Thorson M (1974) Distributed
S adeghi SC, Chacron MJ, Taylor MC, Cullen KE relaxation processes in sensory adaptation. Science
(2007a) Neural variability, detection thresholds, 183:161-172.
and information transmission in the vestibular Tomko DL, Peterka RJ, Schor RH (1981a) Responses
system. J Neurosci 27:771-781. to head tilt in cat eight nerve afferents. Exp Brain
Sadeghi SC, Minor LB, Cullen KE (2007b) Response Res 41:216-221.
of vestibular-nerve afferents to active and passive Tomko D L, Peterka RJ, Schor RH, O'Leary DP
rotations under normal conditions and after (1981b) Response dyn amics of horizontal
unilateral labyrinthectomy. J Neurophysiol canal afferents in barbiturate-anesthetized cats.
97:1503-1514. J Neurophysiol 45:376-396.
Scherer H, Brandt U, Clarke AH, Merbol<l U, Parker Trincker D (1962) The transformation of mechanical
R (1986) European vestibular experiments on the stimulus into nervous excitation by the labyrinthine
Spacelab-1 mission: 3. Caloric nystagmus in micro­ receptors. Soc Exp Biol (Great Britain) 16:
gravity. Exp Brain Res 64:255-263. 289-316.
Scherer H, Clarke AH (1985) The caloric vestibular van Egmond AAJ, Groen, J.J., Jongkees LBW (1949)
reaction in space. Physiological considerations. The mechanics of the semicircular canal. J Physiol
Acta Otolaryngol 100:328-336. 110:1-17.
Schneider LW, Anderson DJ (1976) Transfer Vollrath MA, Eatock RA (2003) Time course and
characteristics of first- and second-order lateral extent of mechanotransducer adaptation in mouse
canal vestibular neurons in gerbil. Brain Res utricular hair cells: comparison with frog saccular
112:61-76. hair cells. J Neurophysiol 90:2676-2689.
Shannon CE, Weaver W (1949) The Mathematical Wersa.11 J (1956) Studies on the structure and innerva­
Theory of Communication. Urbana, IL: University tion of the sensory epithelium of the cristae ampul­
of Illinois Press. laris in the guinea pig. A light and electron
4. PHYSIOLOGY O F T H E VESTI B U LA R ORGANS 115

microscopic investigation. Acta Otolaryngol Suppl Young JH, Anderson DJ (1974) Response patterm of
( Stockh) 126:1-85. primary vestibular neurons to thermal and rota­
Yagi T, Simpson NE, Markham CH (1977) The rela­ tional stimuli. Brain Res 79: 199-212.
tionship of conduction velocity to other physiologi­ Zhao X, Yang H, Yamoah EN, Lundberg YW (2007)
cal properties of the cat's horizontal canal neurons. Gene targeting reveals the role of Oc90 as the
Exp Brain Res 30:587-600. essential organizer of the otoconial organic matrix.
Yang A, Hullar TE (2007) The relationship of semi­ Dev Biol 304:508-524.
circular canal size to vestibular-nerve afferent sensi­
tivity in mammals. J Neurophysiol. 98: 3197-3205.

You might also like