Sie sind auf Seite 1von 11

Cell Calcium 33 (2003) 185–195

Regulation of outer hair cell cytoskeletal stiffness by intracellular Ca2+:


underlying mechanism and implications for cochlear mechanics
Gregory I. Frolenkov a,∗ , Fabio Mammano b , Bechara Kachar a
a Section on Structural Cell Biology, National Institute on Deafness and Other Communication Disorders (NIDCD),
National Institutes of Health (NIH), Bldg. 50, Room 4346, Bethesda, MD 20892-8027, USA
b Istituto Nazionale di Fisica della Materia (INFM) and Venetian Institute of Molecular Medicine (VIMM), Padova, Italy
Received 14 August 2002; received in revised form 25 November 2002; accepted 25 November 2002

Abstract

Two Ca2+ -dependent mechanisms have been proposed to regulate the mechanical properties of outer hair cells (OHCs), the sensory–motor
receptors of the mammalian cochlea. One involves the efferent neurotransmitter, acetylcholine, decreasing OHC axial stiffness. The other
depends on elevation of intracellular free Ca2+ concentration ([Ca2+ ]i ) resulting in OHC elongation, a process known as Ca2+ -dependent
slow motility. Here we provide evidence that both these phenomena share a common mechanism. In whole-cell patch-clamp conditions, a
fast increase of [Ca2+ ]i by UV-photolysis of caged Ca2+ or by extracellular application of Ca2+ -ionophore, ionomycin, produced relatively
slow (time constant ∼20 s) cell elongation. When OHCs were partially collapsed by applying minimal negative pressure through the patch
pipette, elevation of the [Ca2+ ]i up to millimole levels (estimated by Fura-2) was unable to restore the cylindrical shape of the OHC.
Stiffness measurements with vibrating elastic probes showed that the increase of [Ca2+ ]i causes a decrease of OHC axial stiffness, with
time course similar to that of the Ca2+ -dependent elongation, without developing any measurable force. We concluded that, contrary to a
previous proposal, Ca2+ -induced OHC elongation is unlikely to be driven by circumferential contraction of the lateral wall, but is more
likely a passive mechanical reaction of the turgid OHC to Ca2+ -induced decrease of axial stiffness. This may be the key phenomenon for
controlling gain and operating point of the cochlear amplifier.
© 2003 Elsevier Science Ltd. All rights reserved.

Keywords: Cochlear amplifier; Prestin; Olivocohlear bundle; Acetylcholine; Intracellular calcium stores; Slow motility

1. Introduction cochlear partition, which is achieved by keeping the basilar


membrane at the threshold of spontaneous oscillations (for
Cochlear OHCs are responsible for amplification of review, see [9]). Enormous range of input sound pressures
sound-induced vibrations within the organ of Corti (re- of about six orders of magnitude requires especially dedi-
viewed in [1,2]) due to their ability to elongate and shorten cated mechanisms capable of adjusting the gain of cochlear
at acoustic frequencies following the changes of intracellular amplifier. Due to the strategic location of OHCs in the organ
potential (for review, see [3,4]). A unique, ATP-independent of Corti, any change of their mechanical properties, par-
[5], membrane-based [6] form of cellular motility is re- ticularly somatic length and cytoskeletal stiffness, is bound
sponsible for these fast changes of OHC shape, known to profoundly affect cochlear mechanics, for instance by
as electromotility. A recently cloned membrane protein, shifting the operating point of the mechano-sensory appara-
prestin [7], populates the lateral plasma membrane of OHCs tus hosted in the cell stereocilia and/or the local resonance
[8] and serves as a direct electro-mechanical transducer frequency of the cochlear partition.
[4], presumably providing the force required for cochlear In fact, OHCs are the target of an abundant efferent
amplification. innervation originating in the brainstem, whose activation
The main function of the cochlear amplifier is to coun- results in elevation of hearing threshold, i.e. a decrease of
teract intrinsic viscous losses of vibrating energy of the system’s sensitivity (reviewed in [10]). The principal neu-
rotransmitter of this efferent system is acetylcholine (ACh)
(for review see [11]). Whole-cell recordings from isolated
∗ Corresponding author. Tel.: +1-301-402-4211; fax: +1-301-402-1765. OHCs have provided functional evidence for cholinergic
E-mail address: gregory frolenkov@nih.gov (G.I. Frolenkov). receptors localized around the base of the cell, where the

0143-4160/03/$ – see front matter © 2003 Elsevier Science Ltd. All rights reserved.
doi:10.1016/S0143-4160(02)00228-2
186 G.I. Frolenkov et al. / Cell Calcium 33 (2003) 185–195

efferent synapses are located [12]. The action of ACh on (22–24 ◦ C) throughout the experiments. In several control
OHC currents is relatively fast (time scale of tens or hun- experiments, OHCs were placed in Ca2+ -free L-15 medium,
dreds of milliseconds [12,13]), requires extracellular Ca2+ prepared as described above with CaCl2 replaced by 2 mM
[13,14], is accompanied by changes of [Ca2+ ]i [15,16], EGTA.
and is probably mediated by a novel heteromeric receptor,
assembled from ␣9 and ␣10 subunits [17]. 2.2. Patch-clamp recordings
In addition to fast Ca2+ -dependent ionic currents stim-
ulated by ACh, two slower Ca2+ -dependent mechanisms, OHCs were visualized with an inverted microscope (Ax-
operating in a time scale of tens of seconds and regulating iovert 200, Carl Zeiss, Gottingen, Germany) and the follow-
the mechanical properties of cochlear OHCs, have been ing morphological features were used to determine viability:
proposed. On the one hand, ACh would decrease OHC uniform cylindrical shape, basal location of the nucleus,
axial stiffness [18] presumably through Ca2+ -dependent membrane birefringence and intact stereocilia. Patch-clamp
phosphorylation of unspecified cytoskeletal proteins [19] pipettes were filled with an intracellular solution containing
without appreciable effects on OHC electromotile mecha- (in mM): KCl (144), MgCl2 (2.0), EGTA (0.5), Na2 HPO4
nism [20]. On the other hand, elevation of [Ca2+ ]i would (8.0), NaH2 PO4 (2.0), Mg-ATP (2.0), Na-GTP (0.2), ad-
trigger a circumferential contraction of the OHC cortical justed to pH 7.2 with KOH and brought to 325 mOsm
cytoskeleton resulting in axial elongation of the OHC [21]. with d-glucose. Patch-clamp recordings were performed
This elongation, known as “Ca2+ -dependent slow motility”, using an Axopatch 1D amplifier (Axon Instruments, Foster
involves Ca2+ /calmodulin-dependent protein phosphoryla- City, CA). Pipettes for conventional whole-cell recordings
tion [22,23]. were formed on a programmable puller (P87, Sutter Instru-
Here we used patch-clamp recordings, UV photolysis, ments, Novato, CA) from 1.0 mm O.D. borosilicate glass
Ca2+ -imaging, and stiffness measurement techniques to (#30-30-0, FHC, Bowdoinham, ME). Current and voltage
study the properties Ca2+ -induced slow motility and its rela- were sampled at 100 kHz using a standard laboratory in-
tionship with cytoskeletal stiffness. terface (Digidata 1200A, Axon Instruments) controlled by
pClamp 7.0 software (Axon Instruments). The uncompen-
sated pipette resistance was typically 3–5 M when mea-
2. Methods sured in the bath and the access resistance did not exceed
15 M under whole-cell patch-clamp conditions. Potentials
2.1. Cell preparation were corrected off-line for the error due to the access resis-
tance. Junction potential was −4.2 mV, as computed by the
Adult guinea pigs (200–400 g) were sacrificed by suffoca- pClamp 7.0 software based on the given solution compo-
tion with carbon dioxide and decapitated according to NIH sition. This value was rather small and no correction was
Guidelines for Animal Use. The temporal bones were re- applied to the data for liquid-junction potential.
moved from the skull and placed in a modified Leibowitz cell
culture medium (L-15) containing (in mM): NaCl (137), KCl 2.3. Cell loading
(5.4), CaCl2 (1.3), MgCl2 (1.0), Na2 HPO4 (1.0), KH2 PO4
(0.44), MgSO4 (0.81). Osmolarity and pH were adjusted to To monitor changes of [Ca2+ ]i , we used two Ca2+ -sensi-
325 ± 2 mOsm with d-glucose and 7.35 mOsm with NaOH, tive fluorescent indicators (Molecular Probes, Eugene, OR):
respectively. To isolate OHCs, the bulla was opened to ex- Calcium Green-1 for experiments requiring photolysis of
pose the cochlea, and the otic capsula was chipped away caged compounds, and Fura-2, otherwise. These dyes were
with a surgical blade starting from the base. Strips of the loaded into the cell either through the patch pipette or by in-
organ of Corti were dissected from the modiolus with a cubation with their cell-permeable AM-ester derivatives. In
fine needle, transferred with a glass pipette to a 100 ␮l the patch-clamp experiments, the intra-pipette solution was
drop of medium containing 1 mg/ml of collagenase type prepared as described earlier, omitting EGTA, and supple-
IV (Life Technologies, Rockville, MD), and incubated for menting it with 50 ␮M of the cell-impermeable form of one
15–20 min. When experiments required preincubation with of these dyes. In the experiments with free isolated OHCs,
the AM-ester form of Ca2+ -sensitive indicators and/or pho- the acetoxymethyl ester form of the dye (10 mM stock so-
tolabile Ca2+ -chelators (see the description later), collage- lution in DMSO) was dispersed with the equal amount of
nase was added 15–20 min before the end of the incubation 20% DMSO-based solution of Pluronic F-127 (Molecular
period. After incubation, cells were dissociated by gentle Probes), diluted to the final concentration of 10 ␮M with
reflux of the tissue through the needle of a Hamilton sy- L-15, and used to pre-incubate OHCs for 30 min at room
ringe (N. 705, 22 gauge) and allowed to settle on the slide temperature.
for 5–10 min. OHCs were placed in a laminar flow bath The increase of [Ca2+ ]i was induced either by ex-
(100 ␮l), with exchange of L-15 medium (about 5 ml/h) by a tracellular application of the Ca2+ -ionophore ionomycin
pressurized perfusion system (BPS-4; ALA Scientific Instru- (Sigma-Aldrich, St. Louis, MO) or by photolysis of a pho-
ments, Westbury, NY), and maintained at room temperature tolabile Ca2+ chelator, pre-complexed with Ca2+ . In the
G.I. Frolenkov et al. / Cell Calcium 33 (2003) 185–195 187

patch-clamp experiments, NP-EGTA (Molecular Probes) were formed on a scientific grade cooled CCD sensor (Mi-
at concentration of 2 mM was mixed with Ca2+ (1.5 mM), cromax 1300Y, Roper Scientific, Trenton, NJ) using 1.6×
added to the intra-pipette solution without EGTA, and sup- secondary magnification. Sensor’s output was binned 2 × 2,
plemented with 50 ␮M of Calcium Green-1. In other experi- digitized at 12 bit/pixel, recorded to a host PC controlled by
ments, OHCs were incubated with the cell-permeable caged the MetaMorph software, and analyzed off-line.
Ca2+ compound Nitr-5/AM (Calbiochem, San Diego, CA) Before subjecting OHCs to any sort of stimulation, 30
at 25 ␮M concentration for 30 min (at room temperature). control images were acquired to determine baseline. (Here
In the latter case, the chelator bound only the free Ca2+ and thereafter, we shall designate “image” as a single
inside the cell and the exact amount of Ca2+ complexed to non-ratiometric fluorescent image of the cell or the pairs
the photoactive substance remained undetermined. of images taken at different wavelengths when the cell was
loaded with Fura-2.) During the experiment, either the injec-
2.4. Drug delivery tion system was activated to apply ionomycin, or broadband
UV illumination of maximal intensity was delivered for
Stock solution of ionomycin in DMSO (10 mM) was di- 1–3 s to initiate photolysis of caged compounds. To track the
luted in the extracellular solution to a final concentration of effect of stimulation, 120 images were acquired. Exposure
5 or 10 ␮M and was used to fill a puff pipette, prepared sim- times depended on the intensity of fluorescence and were
ilarly to the patch pipette. This pipette was placed near the typically in the range of 200–400 ms for non-ratiometric
basolateral wall of the OHC and pressure (10–15 kPa) was imaging and 200–800 ms for ratiometric imaging. Time de-
applied to its back by a pneumatic injection system (PLI-100, lay for switching filters/attenuations was negligible (∼1 ms)
Medical Systems Corp., Greenvale, NY) gated under soft- compared to typical exposure times. Overall, sampling rate
ware control. Typical drug delivery time was ≤30 ms, as de- was in the range from 1 to 2.5 images per second.
termined by positioning the patch pipette in front of the puff When the cell was loaded with Fura-2, fluorescence
pipette and monitoring the change in junction potential. was excited at the standard wavelength pair of 340 and
380 nm to estimate [Ca2+ ]i from consecutive pair ratios
2.5. Ca2+ imaging and photolysis of caged compounds [24]. Background-subtracted fluorescence ratio images were
calibrated using a calcium calibration buffer kit (C-3722,
Light from a 175 W stabilized Xenon arc source (Lambda Molecular Probes). Background fluorescence was calcu-
DG-4, Sutter instruments) was coupled via a liquid light lated for each image individually as the average intensity of
guide to the epifluorescence port of our Axiovert 200 fluorescence in a manually selected reference region, com-
microscope, equipped with an oil-immersion UV-passing mon to the whole sequence, devoid of any obvious cellular
objective (Carl Zeiss, F-Fluar, 40×, 1.3 N.A.). The follow- structures.
ing excitation filters were mounted into the light source: To estimate the overall changes of [Ca2+ ]i in the OHC,
(1) a band-pass filter used to excite Calcium Green-1 we constructed a region of interest covering the whole
(HQ480/40×, Chroma Technology, Brattleboro, VT); (2) a cell body for each image in the sequence. For ratiometric
broad-band UV filter (330WB80, Omega Optical, Brattle- measurements, the average ratio of 340/380 nm fluorescent
boro, VT) for uncaging photolabile Ca2+ chelators; (3) a signals from this region was used to estimate changes of
band-pass filters to excite Fura-2 (340HT15 and 380HT15, average [Ca2+ ]i in OHCs. When the single-wavelength
Omega Optical). The light source was pre-programmed to indicator was used, average fluorescence signals from
deliver 0, 25, 50, and 75% light attenuation at each filter this region of interest was computed as ratio F /F =
position. Additional attenuation was provided to the 480 nm [F (t) − F (0)]/F (0), where t is time, F(t) is fluorescence
(Calcium Green-1) excitation channel by a neutral density following a stimulus that causes Ca2+ elevation within the
filter. Switching of the filter positions and attenuation steps cell, and F(0) is pre-stimulus (baseline) fluorescence. This
was controlled by the acquisition software (MetaMorph, procedure monitors changes of [Ca2+ ]i correctly only if the
Universal Imaging Co., Downingtown, PA). The illumi- cell does not change the shape. In our experiments, OHCs
nation intensity was adjusted to avoid phototoxicity by elongated, the area of the corresponding region of interest
reducing dye photo-bleaching rates to ≤0.1%/s. Decompo- increased, and changes of the [Ca2+ ]i were thus generally
sition of the caged compounds was minimized by inserting underestimated. Aware of this potential pitfall we refrained
a green interference filter into microscope condenser to cut from drawing quantitative conclusions from non-ratiometric
off the wavelengths in the near-UV range from bright-field [Ca2+ ]i measurements.
illumination.
The reflector turret of the microscope was equipped 2.6. Stiffness measurement
with a special dichroic mirror for simultaneous excitation
in the UV and 480 nm spectral regions (#74100, Chroma Elastic probes were pulled from the same 1.0 mm O.D.
Technology) and a band-pass observation filter optimized borosilicate glass used for patch pipette manufacturing. The
for the Ca2+ -selective dyes Calcium Green-1 and Fura-2 tapered fiber of each probe was 10–15 mm in length and
(HQ535/50m, Chroma Technology). Fluorescence images ∼0.5 ␮m in tip diameter. We did not measure the actual
188 G.I. Frolenkov et al. / Cell Calcium 33 (2003) 185–195

Fig. 1. Measurement of OHC axial stiffness. (A) Positioning of the elastic probe (P) on the cuticular plate (CP) of the OHC; St, stereocilia. (B) Apparent
ratio of cell/probe stiffness (see Section 2) estimated at four different positions of the probe, starting from a point above the cuticular plate and moving
towards the cell (arrows). Shaded region indicates the position of the cell edge, where the probe contacted the cuticular plate. Inset—movements of the
elastic probe unattached to the cell (unloaded) and maximally pushed against the cuticular plate (loaded).

stiffness of the probe, but our manufacturing protocol was ensuring that recording conditions or cell stiffness did not
optimized to ensure that it was comparable to the OHC ax- change spontaneously.
ial stiffness. For all OHCs included in the analysis, this was Assuming the probe and the cell as two sequentially con-
ascertained by pushing the probe against the cell’s cuticular nected elastic elements with stiffness Kp and Kc , respec-
plate and confirming that cell compression and probe de- tively, the stiffness ratio could be estimated as: Kc /Kp =
flection were comparable. To measure relative stiffness, the Au /Al − 1, where Au and Al are unloaded and loaded vi-
probe was attached to a vibrating piezo actuator mounted on bration amplitude, respectively. Since only Kc is expected
a 3-axis micromanipulator (Narishige, Tokyo, Japan). The to change after OHC stimulation, percent changes of Kc /Kp
amplitude of vibration was <1.5 ␮m to minimize poten- simply reflect underlying percent changes of Kc .
tial traumatic effects of external mechanical stress on the
cell. 2.7. Measurements of cell and probe movement
The isolated OHCs were held by the patch pipette, located
just above the nucleus, by applying gentle suction without Measurements of cell and/or elastic probe motion were
breaking the patch. The probe was oriented parallel to the performed as described in [25]. Briefly, motion was recorded
OHC cuticular plate and approximately perpendicular to the with a video camera interfacing with an inverted microscope
long axis of the cell (Fig. 1A). OHCs with tilted cuticular equipped with Differential Interference Contrast (DIC) op-
plates were not included in this study. Care was taken to tics to an optical disk recorder (Panasonic TQ-3031F). Dig-
keep the probe away from the hair bundle, as far as possible, itized images were analyzed off-line with the MetaMorph
without compromising its ability to produce axial compres- image-processing system. For motion quantification, a mea-
sion of the cell and without tilting the cuticular plate. This suring rectangle ranging in length from 5 to 20 ␮m and
arrangement provides a predominant compression of the cell composed of 3–15 rows of pixels was positioned across the
body in axial direction even for curved OHCs. Despite some moving edge of the cell or the probe. The average intensity
possible radial component, we will designate stiffness esti- profile across the edge of the object was calculated and the
mated in this arrangement as “axial stiffness”. number of points in the profile was increased 10 times by cu-
Sinusoidal command voltages with a frequency of 1–3 Hz bic spline interpolation. Motion of the object was calculated
drove the vibrating motion of the probe along the OHC axis from the frame-by-frame shift (computed by a least-square
in such a way that it would alternatively compress and relax procedure) in the interpolated intensity profiles. The sensi-
the cell. For each cell, the motion of the freestanding probe tivity of the measurement was ∼0.02 ␮m, as previously de-
(unloaded vibration) was recorded first (Fig. 1B, inset). The termined [25].
probe was then carefully advanced to compress the cell, until
its vibration amplitude stopped decreasing, indicating full 2.8. Statistical analysis
contact. Under these conditions, the apparent cell stiffness
reached a plateau value (Fig. 1B). The effects of [Ca2+ ]i All values are given as means ± S.E.M. unless otherwise
changes on the OHC stiffness were monitored by measuring stated. Statistical significance was estimated using Student’s
vibration amplitude around this position. Before stimulating t-test at P = 0.05 level. Curves were fitted to data by a
the cell, a baseline record of 15–40 s duration was obtained, Levenberg–Marquardt algorithm using Origin 6.0 software
G.I. Frolenkov et al. / Cell Calcium 33 (2003) 185–195 189

(Microcal Software, Northampton, MA). This software es- 0.8–4.5 ␮M (n = 10 cells). In 12 out of 15 cells maintained
timated also the standard errors of fitting parameters. under whole-cell patch-clamp conditions, ionomycin also
elicited an outward current of 40–280 pA when measured
at holding potentials in the range from −70 to −55 mV
3. Results (Fig. 2C). This current, which has been previously ob-
served in OHCs [12,20,26], reflected likely the activation
3.1. Measurement of Ca2+ -induced elongation of OHCs of Ca2+ -dependent K+ -channels. Although the intracellu-
in whole-cell patch-clamp conditions lar potential was fixed, the ionomycin-induced increase of
[Ca2+ ]i was followed by a relatively slow elongation of the
Fig. 2A shows a typical response of a relatively long OHC. In a sample experiment shown in Fig. 2C, the latency
OHC to the pressure application of the Ca2+ -ionophore ion- of this elongation was about 10 s and the time constant was
omycin. When applied at 5 ␮M concentration, ionomycin 15.8 ± 0.9 s. Cell elongation and [Ca2+ ]i increase were sup-
produced a dramatic increase of the [Ca2+ ]i from base- pressed after more than 15 min of incubation in Ca2+ -free
line values of 145–220 nM (typical of isolated OHCs) to extracellular medium (data not shown). Therefore, we shall

Fig. 2. Ca2+ -induced OHC currents and motile responses in whole-cell patch-clamp conditions. (A, C) Responses of an OHC to the application of
ionomycin (10 ␮M). (B, D) Responses of another OHC to increase of the [Ca2+ ]i produced by UV-photolysis of the photolabile Ca2+ -chelator NP-EGTA
(2 mM) pre-loaded with Ca2+ (1.5 mM) and introduced into the cell through the patch pipette. (A, B) Pairs of OHC images showing the distribution
of the [Ca2+ ]i before (left images in the pair) and after (right images) stimulation. [Ca2+ ]i changes were estimated with the ratiometric fluorescent
Ca2+ -indicator Fura-2 in A, and the single-wavelength indicator Calcium Green-1 in B. (C, D) From top to bottom: changes of the total Ca2+ -dependent
fluorescent signal emitted from OHCs; relative changes of OHC length; whole-cell current. The timing of ionomycin application is indicated by a
horizontal bar (C), the timing of UV-illumination is indicated by two arrows (begin–end, D). The measurement procedures are described in Section 2.
OHC in A and C was held at −48 mV, OHC in B and D was held at −53 mV.
190 G.I. Frolenkov et al. / Cell Calcium 33 (2003) 185–195

hereafter attribute the ionomycin-induced OHC elongation shown). This current was adapting even if [Ca2+ ]i remained
to the “Ca2+ -induced slow motility” observed previously elevated (Fig. 2D). In contrast to the relatively fast changes
in free isolated unpatched OHCs [21]. of [Ca2+ ]i and whole-cell current, changes of OHC length
The use of UV-photolysis affords an alternative method occurred with slower time constant (∼10 s for the cell il-
to gate the increase of the [Ca2+ ]i with better temporal lustrated on Fig. 2D), similar to the ionomycin-induced
resolution. Photolysis of the UV-sensitive Ca2+ chelator, elongation (Fig. 2A). In four out of five cells, the elonga-
NP-EGTA, pre-complexed with Ca2+ and introduced into tion of the OHC initiated by UV-photolysis was bi-phasic
the cell through the patch pipette, produced fast and dis- (Fig. 2D) without obvious correlation with the time course
tinct increase of [Ca2+ ]i in OHCs (Fig. 2B). However, of the fluorescence change.
we were unable to estimate the absolute values of [Ca2+ ]i
changes because of the difficulties involved in the simulta- 3.2. Comparing the time course of Ca2+ -induced OHC
neous usage of the ratiometric Ca2+ -indicator Fura-2 with elongation in free cells and under patch-clamp conditions
photolabile compounds [27]. The fluorescence of OHCs
loaded with Calcium Green-1, a non-ratiometric indicator In a separate set of experiments, we studied Ca2+ -induced
excited in the visible region of the spectrum, increased elongation in free (unpatched) OHCs. To monitor [Ca2+ ]i
only by 5.4 ± 0.3% after photolysis (n = 4). The amount changes evoked by UV-photolysis, OHCs were loaded with
of Ca2+ released by UV illumination depends on the in- the cell permeant forms of Nitr-5 and Calcium Green-1 (see
tracellular concentration of the chelator. After establish- Section 2). Fig. 3 presents average responses obtained in
ing whole-cell configuration, the chelator slowly diffused these conditions and compares them with the correspond-
into the cell and repeated application of UV illumina- ing results of patch-clamp experiments. UV-activated Ca2+
tion produced progressively larger changes of fluorescent photo-release was less effective in free unpatched OHCs than
signal, reaching a plateau after 15–30 min of whole-cell in patch-clamped OHCs, as expected given that the mem-
recording. The Ca2+ -induced elongation of OHCs could brane permeant form of the photolabile Ca2+ chelator may
be observed during the first 10–15 min of whole-cell only complex with the limited amount of free intracellular
patch-clamp recording, suggesting that it requires some Ca2+ (less or about 200 nM, compared to 1.5 mM of the
soluble intracellular components, which were inevitably patch pipette mixture; see Section 2). Nonetheless, measur-
washing out in the whole-cell recording. For these reasons, able elongations of the OHC were still evoked under these
data collection was terminated after 15 min of whole-cell unfavorable conditions and were appreciably faster than in
recording. the patch-clamp recordings. Furthermore, cell length in-
The increase of the [Ca2+ ]i by UV-photolysis evoked a creased with short latency, typically shorter than the duration
noticeable outward current (50 pA for the cell illustrated in of the UV-irradiation in these experiments (<3 s). Finally,
Fig. 2D, 30–110 pA for four out of five OHCs tested). This the time constant of OHC elongation (6.5 ± 0.7 s, Fig. 3A)
current was maximal at holding potentials around −20 mV, was significantly faster than that observed in patch-clamp
as expected for a Ca2+ -activated K+ -current (data not experiments (18 ± 1 s, Fig. 3B, P < 0.0001). Since this fast

Fig. 3. Ca2+ -dependent elongation is faster in isolated unpatched (A) vs. patched (B) OHCs. Rise of the [Ca2+ ]i in these cells was produced by
UV-photolysis of photolabile Ca2+ -chelators, introduced into the cell either by pre-incubation with the AM-ester derivative (Nitr-5 in A) or through the
patch pipette (NP-EGTA in B). Timing of UV illumination is indicated by two arrows (begin–end). Top records on each panel show the relative changes
of the Calcium Green-1 fluorescence signal averaged over the whole cell body. Bottom records illustrate relative changes of cell length. Each record is
an average of the responses observed in n = 8 (A) and n = 4 (B) cells, respectively. Error bars on the left side of each record indicate S.E.M. values.
G.I. Frolenkov et al. / Cell Calcium 33 (2003) 185–195 191

phase of the Ca2+ -dependent OHC elongation was missing 230 nM to 4.5 ␮M. [Ca2+ ]i changes were followed by a slow
when the intracellular potential was held constant, we con- increase of probe vibration amplitude (Fig. 5C), indicating
clude that it resulted from a voltage-driven elongation of a twofold decrease of OHC axial stiffness (or more), only
OHC activated by Ca2+ -induced hyperpolarization, whereas partially reversible after the end of application (Fig. 5D). In
the slow Ca2+ -induced component of the elongation was all five cells tested, ionomycin produced a decrease of ax-
probably too small to be detected under these experimental ial stiffness (range of changes: 16–80%). The time course
conditions. of these changes was similar to that of ionomycin-induced
OHC elongation (Fig. 2A). Ionomycin-induced changes of
3.3. Ca2+ -induced changes of OHC turgor OHC axial stiffness were inhibited either after more 15 min
of incubation in Ca2+ -free extracellular medium or dur-
The cylindrical shape of OHCs results from positive in- ing the latest stages of patch-clamp recording (>15 min af-
tracellular pressure (cell turgor [28]). Thus, if cell elonga- ter the establishment of whole-cell configuration), when the
tion were actually driven by circumferential contraction of Ca2+ -induced elongation could not be evoked (data not
the lateral wall, as proposed [21], the elevation of [Ca2+ ]i shown).
should, in principle, promote restoration the shape of an To address the question whether Ca2+ -induced elongation
OHC partially collapsed by applying small negative pres- of the OHC could produce any mechanical work against an
sure via the patch pipette. This was never observed in any external load, we measured the resting length of the cell un-
of n = 4 cells tested (Fig. 4A). Instead, OHCs often tended der static compression by an elastic probe (Fig. 6). When
to lose their cylindrical shape after ionomycin (Fig. 4B). the stiffness of the elastic probe was comparable to the OHC
Changes in cell turgor were most prominent when [Ca2+ ]i axial stiffness, ionomycin application invariably resulted in
increased above about 5 ␮M as determined by Fura-2 fluo- further compression of the loaded cell, indicating that me-
rescence. Cell shape recovered only partially after washout chanical work was done against the cell (Fig. 6A). When
(Fig. 4B). the cell was either not loaded, or loaded by a probe having
comparatively negligible stiffness, a clear ionomycin-evoked
3.4. Ca2+ -activated decrease of OHC axial stiffness elongation was observed (Fig. 6A). These responses were
dramatically different from electrically evoked motile re-
OHC axial stiffness was monitored by pushing a vibrating sponses observed in the same cells when the intracellular
elastic probe against the cuticular plate (Fig. 5A). In a typi- potential was altered under voltage-clamp. In particular, the
cal experiment, illustrated in Fig. 5B, the amplitude of probe electrically-evoked elongation evoked by hyperpolarization
vibration decreased by about five times after applying me- easily deflected also those probes that possessed a stiffness
chanical force to the cell, suggesting that the cell/probe stiff- 2–3 times larger than the cell stiffness, indicating that me-
ness ratio was close to 4. In the same experiment, application chanical work was done against the elastic load of the (ex-
of ionomycin increased the [Ca2+ ]i from a basal value of ternal) probe (Fig. 6B).

Fig. 4. Ionomycin-induced decrease of OHC turgor. (A) Fluorescent images of an OHC before (left) and 40 s after transient (20 s duration) application of
10 ␮M of ionomycin (right). The cell was slightly deflated by applying negative pressure to the patch pipette. Intensity of fluorescence monitors [Ca2+ ]i
in arbitrary units. OHC was in whole-cell patch-clamp at holding potential of −53 mV. (B) DIC images of another OHC with apparently normal turgor
before ionomycin application (left), 40 s following transient application of 10 mM of ionomycin, and 150 s thereafter (right). Arrows point to the locations
where OHC lateral wall curved inward, indicating that the cell started to lose their turgor. Holding potential was −20 mV.
192 G.I. Frolenkov et al. / Cell Calcium 33 (2003) 185–195

Fig. 5. Ca2+ reduces OHC axial stiffness. Panel A shows a differential interference contrast (DIC) image of an OHC (top) and the movement of the
elastic probe (bottom), which was vibrating in close proximity to the cuticular plate without touching the cell. Panel B shows DIC image (top, left)
and Fura-2 ratio image (top, right) of the same cell compressed longitudinally by the probe. Bottom record illustrates movement of the vibrating elastic
probe. Panel C is the same as B, 15 s following the application of ionomycin (10 ␮M). (D) Time course of the stiffness changes of this cell. Each point
represents an average of five stiffness measurements calculated for five consecutive vibration cycles. S.E.M. values are indicated. Horizontal bar marks
timing of ionomycin application.

Fig. 6. Ionomycin (A) and hyperpolarization (B) produce significantly different mechanical forces. Top records illustrate ionomycin and hyperpolarization
induced changes of OHC length without any mechanical load attached to the cell. Bottom records show the length responses of the cell compressed by
the elastic probe. Data on panel A were obtained in two different cells, data on panel B were recorded from the same cell. Horizontal bars mark either
timing of ionomycin application (A) or negative voltage command (B).

4. Discussion cochlear potentials were observed in vivo: a “fast” effect,


with a time constants of about 100 ms, and a “slow” one,
Efferent-mediated sound-induced changes of cochlear with characteristic time constant of 25–50 s [32]. Both ef-
function require 100–200 ms to develop in anesthetized cats fects are initiated by the entry of Ca2+ into OHCs through
[29] and at least 13–15 ms to develop in awake humans and acetylcholine receptor [33] and accompanied by changes
guinea pigs [30,31]. Two distinct modulatory effects of the of the sound responses of the basilar membrane [34]. At
electrical stimulation of the olivo-cochlear bundle on the the level of OHCs, “fast” effects have been tentatively
G.I. Frolenkov et al. / Cell Calcium 33 (2003) 185–195 193

attributed to ACh-induced ionic currents, while “slow” ef- interpreted as a result of increased cell rigidity [21]. How-
fects may be associated with slow Ca2+ -induced changes of ever, this interpretation is challenged by others’, as well as
OHC shape and/or stiffness [18]. Here we show that “slow” our own results. First, the OHCs show prominent regulatory
Ca2+ -dependent elongation of isolated OHCs is primarily volume decrease (RVD) in response to hypo-osmotic stim-
an effect of “slow” Ca2+ -induced decrease of axial stiffness. ulation [38]. Any Ca2+ -dependent modulation of the RVD
mechanisms may affect steady-state volume measurements
4.1. OHC hyperpolarization has a minor effect in like those reported by Dulon et al. [21]. Second, as shown
Ca2+ -dependent slow motility here, Ca2+ -induced elongation in free unpatched OHCs has
at most a minor component associated with cell hyperpo-
After the original experiments of Dulon et al. [21], larization (Fig. 3). Hyperpolarization, in turn, increases the
it was pointed out [22] that the generalized increase axial stiffness of OHC [39,40], partially counteracting the
of [Ca2+ ]i produced by Ca2+ -ionophores may induce primary effect of Ca2+ . Finally, large changes of [Ca2+ ]i
Ca2+ -dependent phosphorylation of numerous proteins, in- may disturb OHC turgor, decreasing also the apparent vol-
cluding Ca2+ -dependent modulation of the various plasma ume changes in response to hypo-osmotic stimulation. In
membrane ion channels. Changes of ionic conductance the present experiments, we performed direct measurements
likely affect intracellular potential, and in turn may result of the OHC axial stiffness with the elastic probe technique
in contraction/elongation of the cell because OHC length is and showed that it decreases in response to increase of
also a function of intracellular potential [35]. This can be the [Ca2+ ]i .
described as “activating fast OHC electromotility slowly”. Our stiffness measurement technique detects elastic defor-
The involvement of a sub-class of K+ -conductances in mations of the cell, imposed by loading the cuticular plate,
Ca2+ -induced OHC elongation has been excluded based on which in turn act as a “piston”. Elastic forces contributing
the fact that length changes were not inhibited by tetraethy- to these deformations may include: (1) axial elasticity of the
lammonium ions [21]. However, other conductances could, mechanical skeleton of the cell; (2) some potential compo-
in principle, effect these changes. Here we have shown nent related to the bending elasticity of the cell skeleton;
that Ca2+ -induced elongation of OHCs can be recorded (3) additional forces resulted from re-distribution of external
in patch-clamp conditions (Fig. 2), when the intracellular mechanical load by intracellular pressure. Of these compo-
potential is fixed. In other words, our data indicate that nents, bending stiffness was probably irrelevant in our ex-
the effect of Ca2+ -ionophores is not mediated by a slow periments, because we observed Ca2+ -induced decrease of
hyperpolarization of the cell. However, the fast component cell stiffness even in perfectly straight OHCs. Axial stiffness
of the fourfold smaller elongation observed in free un- of OHC cytoskeleton is the most likely target for [Ca2+ ]i ac-
patched OHCs probably results from cell hyperpolarization tion. However, externally driven cell compression certainly
mediated by Ca2+ -activated K+ -channels (Fig. 3). increases intracellular pressure providing an additional elas-
The results of our UV-photolysis experiments with photo- tic force, which may depend also on the stiffness of non-axial
labile Ca2+ -chelators, following relatively short (5–15 min) components. Therefore, the Ca2+ -induced decrease of OHC
loading of the cell via patch pipettes, indicate that stiffness observed in our experiments may include also con-
Ca2+ -induced OHC elongation occurs primarily in response tributions from other non-radial mechanical elements of the
to changes of Ca2+ concentration in the cytoplasm. How- cell. In any case, such decrease is hardly compatible with
ever, several intracellular membranes, including those of previously proposed circumferential contraction of the cor-
sarcoplasmic reticulum [36] and mitochondria [37] become tical cytoskeleton [21].
permeable to Ca2+ after application of Ca2+ -ionophores. A Ca2+ -dependent decrease of OHC axial stiffness was
Thus, the ionomycin experiments may include contributions also observed following the application of acetylcholine to
from different intracellular compartments. In OHCs, based the basal pole of the OHC [18]. In isolated unloaded OHCs,
on our data with ACh [16,20], we expect Ca2+ -induced axial stiffness decrease is thought to be responsible for the
Ca2+ -release from intracellular stores to play an essential observed increase of electromotile responses [18]. This can
role in the process. be easily explained if Ca2+ affects the cytoskeletal stiff-
ness more than the effective stiffness of the prestin-based
4.2. Axial stiffness of OHC is regulated by [Ca2+ ]i motors. We showed previously that ionomycin-induced
rise of [Ca2+ ]i mimics the effects of acetylcholine on the
Dulon et al. [21] also suggested that Ca2+ -induced elon- electrically-evoked motile responses of OHCs and on the
gation of the OHC results from a circumferential contraction OHC voltage-dependent capacitance [20], suggesting some
of the cortical cytoskeleton, which, under constant volume common mechanism. Here we round up those observations
conditions, produced cell elongation due to incompressibil- reporting that the effects of the [Ca2+ ]i increase are, indeed,
ity of the cytosol. This argument was based on the obser- associated with a decrease of OHC axial stiffness.
vation that the increase of [Ca2+ ]i decreased the volume The biochemical pathway targeting OHC axial stiffness
responses of the OHCs to the application of hypotonic extra- either after acetylcholine application or following the in-
cellular medium. This decrease of the volume responses was crease of [Ca2+ ]i , remains obscure. What is known is that
194 G.I. Frolenkov et al. / Cell Calcium 33 (2003) 185–195

several inhibitors of calmodulin and calmodulin-dependent [11] M. Eybalin, Neurotransmitters and neuromodulators of the mam-
kinases affect Ca2+ -induced elongation of OHCs [21–23] malian cochlea, Physiol. Rev. 73 (1993) 309–373.
[12] G.D. Housley, J.F. Ashmore, Direct measurement of the action of
and acetylcholine-induced decrease of OHC stiffness [19].
acetylcholine on isolated outer hair cells of the guinea pig cochlea,
The exact cellular mechanism, however, may be fairly com- Proc. R Soc. Lond. B. Biol. Sci. 244 (1991) 161–167.
plicated including regulation of OHC cortical cytoskeleton [13] C. Blanchet, C. Erostegui, M. Sugasawa, D. Dulon, Acetylcholine-
via Rho GTPases [41]. induced potassium current of guinea pig outer hair cells: its
The apparent difference between the effects of ionomycin/ dependence on a calcium influx through nicotinic-like receptors,
J. Neurosci. 16 (1996) 2574–2584.
UV-photolysis and that of acetylcholine is the amount
[14] M.G. Evans, Acetylcholine activates two currents in guinea-pig outer
of [Ca2+ ]i increase involved. Ca2+ -imaging experiments hair cells, J. Physiol. (Lond.) 491 (1996) 563–578.
rarely showed a significant increase of [Ca2+ ]i induced [15] T. Doi, H. Ohmori, Acetylcholine increases intracellular Ca2+ con-
by acetylcholine [15,16]. However, there is indirect evi- centration and hyperpolarizes the guinea-pig outer hair cell, Hear.
dence that acetylcholine may activate spatially localized Res. 67 (1993) 179–188.
Ca2+ -induced Ca2+ release in close vicinity to the plasma [16] M.G. Evans, L. Lagostena, P. Darbon, F. Mammano, Cholinergic
control of membrane conductance and intracellular free Ca2+ in
membrane [16,33], which could not be detected due to lim-
outer hair cells of the guinea pig cochlea, Cell Calcium 28 (2000)
ited resolving power of fluorescence microscopy. Therefore, 195–203.
it is reasonable to conclude that Ca2+ -induced decrease of [17] A.B. Elgoyhen, D.E. Vetter, E. Katz, C.V. Rothlin, S.F. Heinemann,
axial stiffness is a characteristic response of OHCs, which J. Boulter, Alpha10: a determinant of nicotinic cholinergic receptor
evolved as an essential component of the efferent-based function in mammalian vestibular and cochlear mechanosensory hair
cells, Proc. Natl. Acad. Sci. U.S.A. 98 (2001) 3501–3506.
system controlling the mechanical properties of the organ
[18] P. Dallos, D.Z.Z. He, I. Sziklai, S. Metha, B.N. Evans, Acetylcholine,
of Corti. Ca2+ -induced elongation of the OHC is nothing outer hair cell electromotility, and the cochlear amplifier, J. Neurosci.
but a by-product of this active regulation and represents 17 (1997) 2212–2226.
a passive reaction of the turgid cell to a decrease of axial [19] I. Sziklai, M. Szonyi, P. Dallos, Phosphorylation mediates the in-
stiffness. fluence of acetylcholine upon outer hair cell electromotility, Acta
Otolaryngol. 121 (2001) 153–156.
[20] G.I. Frolenkov, F. Mammano, I.A. Belyantseva, D. Coling, B. Kachar,
Two distinct Ca(2+)-dependent signaling pathways regulate the mo-
Acknowledgements tor output of cochlear outer hair cells, J. Neurosci. 20 (2000) 5940–
5948.
We thank David Robinson, Mark Ospeck, Claudia Racca, [21] D. Dulon, G. Zajic, J. Schacht, Increasing intracellular free calcium
and Mark Schneider for critical review of the manuscript. induces circumferential contractions in isolated cochlear outer hair
cells, J. Neurosci. 10 (1990) 1388–1397.
[22] B. Puschner, J. Schacht, Calmodulin-dependent protein kinases me-
diate calcium-induced slow motility of mammalian outer hair cells,
References Hear. Res. 110 (1997) 251–258.
[23] D.E. Coling, S. Bartolami, D. Rhee, T. Neelands, Inhibition of
calcium-dependent motility of cochlear outer hair cells by the protein
[1] J.F. Ashmore, F. Mammano, Can you still see the cochlea for the
kinase inhibitor, ML-9, Hear. Res. 115 (1998) 175–183.
molecules? Curr. Opin. Neurobiol. 11 (2001) 449–454.
[2] L. Robles, M.A. Ruggero, Mechanics of the mammalian cochlea, [24] G. Grynkiewicz, M. Poenie, R.Y. Tsien, A new generation of Ca2+
Physiol. Rev. 81 (2001) 1305–1352. indicators with greatly improved fluorescence properties, J. Biol.
[3] G.I. Frolenkov, M. Atzori, F. Kalinec, F. Mammano, B. Kachar, The Chem. 260 (1985) 3440–3450.
membrane-based mechanism of cell motility in cochlear outer hair [25] G.I. Frolenkov, F. Kalinec, G.A. Tavartkiladze, B. Kachar, Cochlear
cells, Mol. Biol. Cell 9 (1998) 1961–1968. outer hair cell bending in an external electric field, Biophys. J. 73
[4] P. Dallos, B. Fakler, Prestin, a new type of motor protein, Nat. Rev. (1997) 1665–1672.
Mol. Cell. Biol. 3 (2002) 104–111. [26] T. Yamamoto, S. Kakehata, T. Yamada, T. Saito, H. Saito, N. Akaike,
[5] B. Kachar, W.E. Brownell, R. Altschuler, J. Fex, Electrokinetic Effects of potassium channel blockers on the acetylcholine-induced
shape changes of cochlear outer hair cells, Nature 322 (1986) 365– currents in dissociated outer hair cells of guinea pig cochlea, Neu-
368. rosci. Lett. 236 (1997) 79–82.
[6] F. Kalinec, M.C. Holley, K.H. Iwasa, D.J. Lim, B. Kachar, A [27] R.S. Zucker, Effects of photolabile calcium chelators on fluorescent
membrane-based force generation mechanism in auditory sensory calcium indicators, Cell Calcium 13 (1992) 29–40.
cells, Proc. Natl. Acad. Sci. U.S.A. 89 (1992) 8671–8675. [28] M.E. Chertoff, W.E. Brownell, Characterization of cochlear outer
[7] J. Zheng, W. Shen, D.Z. He, K.B. Long, L.D. Madison, P. Dallos, hair cell turgor, Am. J. Physiol. 266 (1994) C467–C479.
Prestin is the motor protein of cochlear outer hair cells, Nature 405 [29] E.H. Warren III, M.C. Liberman, Effects of contralateral sound
(2000) 149–155. on auditory-nerve responses. I. Contributions of cochlear efferents,
[8] I.A. Belyantseva, H.J. Adler, R. Curi, G.I. Frolenkov, B. Kachar, Hear. Res. 37 (1989) 89–104.
Expression and localization of prestin and the sugar transporter [30] G.A. Tavartkiladze, G.I. Frolenkov, S.V. Artamasov, Ipsilateral sup-
GLUT-5 during development of electromotility in cochlear outer hair pression of transient evoked otoacoustic emission: role of the medial
cells, J. Neurosci. 20 (2000) RC116. olivocochlear system, Acta Otolaryngol. 116 (1996) 213–218.
[9] R. Nobili, F. Mammano, J. Ashmore, How well do we understand [31] D.L. da Costa, A. Chibois, J.P. Erre, C. Blanchet, R.C. de Sauvage,
the cochlea? Trends Neurosci. 21 (1998) 159–167. J.M. Aran, Fast, slow, and steady-state effects of contralateral acous-
[10] W.B. Warr, Organization of olivocochlear efferent systems in mam- tic activation of the medial olivocochlear efferent system in awake
mals, in: R.R. Fay, A.N. Popper, D.B. Webster (Eds.), Auditory guinea pigs: action of gentamicin, J. Neurophysiol. 78 (1997) 1826–
Central Nervous System, New York, Springer, 1992, pp. 410–448. 1836.
G.I. Frolenkov et al. / Cell Calcium 33 (2003) 185–195 195

[32] T.S. Sridhar, M.C. Liberman, M.C. Brown, W.F. Sewell, A novel [37] R.F. Kauffman, R.W. Taylor, D.R. Pfeiffer, Cation transport and
cholinergic slow effect of efferent stimulation on cochlear potentials specificity of ionomycin. Comparison with ionophore A23187 in rat
in the guinea pig, J. Neurosci. 15 (1995) 3667–3678. liver mitochondria, J. Biol. Chem. 255 (1980) 2735–2739.
[33] T.S. Sridhar, M.C. Brown, W.F. Sewell, Unique postsynaptic signaling [38] J.R. Crist, M. Fallon, R.P. Bobbin, Volume regulation in cochlear
at the hair cell efferent synapse permits calcium to evoke changes outer hair cells, Hear. Res. 69 (1993) 194–198.
on two time scales, J. Neurosci. 17 (1997) 428–437. [39] G.I. Frolenkov, I.A. Belyantseva, B. Kachar, Electromotility influ-
[34] N. Cooper, J.J. Guinan, Fast and slow olivocochlear efferent effects ences the axial stiffness of the outer hair cells. Assoc. Res. Oto-
on basilar membrane motion involve different mechanisms, Assoc. laryngol. (1998) Abstract #254.
Res. Otolaryngol. (2002) Abstract #312. [40] D.Z. He, P. Dallos, Somatic stiffness of cochlear outer hair cells is
[35] J.F. Ashmore, A fast motile response in guinea-pig outer hair cells: voltage-dependent, Proc. Natl. Acad. Sci. U.S.A. 96 (1999) 8223–
the cellular basis of the cochlear amplifier, J. Physiol. (Lond.) 388 8228.
(1987) 323–347. [41] F. Kalinec, M. Zhang, R. Urrutia, G. Kalinec, Rho GTPases mediate
[36] T.J. Beeler, I. Jona, A. Martonosi, The effect of ionomycin on the regulation of cochlear outer hair cell motility by acetylcholine,
calcium fluxes in sarcoplasmic reticulum vesicles and liposomes, J. J. Biol. Chem. 275 (2000) 28000–28005.
Biol. Chem. 254 (1979) 6229–6231.

Das könnte Ihnen auch gefallen