Sie sind auf Seite 1von 9

Review

TRENDS in Neurosciences

Vol.29 No.10

Acid-sensing ion channels: advances,


questions and therapeutic
opportunities
John A. Wemmie1,2, Margaret P. Price3 and Michael J. Welsh3,4,5
1

Department of Psychiatry, Roy J. and Lucille A. Carver College of Medicine, University of Iowa, Iowa City, IA 52242, USA
Department of Veterans Affairs Medical Center, Iowa City, IA 52242, USA
3
Department of Internal Medicine, Roy J. and Lucille A. Carver College of Medicine, University of Iowa, Iowa City, IA 52242, USA
4
Department of Physiology and Biophysics, Roy J. and Lucille A. Carver College of Medicine, University of Iowa, Iowa City,
IA 52242, USA
5
Howard Hughes Medical Institute, Roy J. and Lucille A. Carver College of Medicine, University of Iowa, Iowa City, IA 52242, USA
2

Extracellular acid can have important effects on neuron


function. In central and peripheral neurons, acid-sensing
ion channels (ASICs) have emerged as key receptors for
extracellular protons, and recent studies suggest diverse
roles for these channels in the pathophysiology of pain,
ischemic stroke and psychiatric disease. ASICs have also
been implicated in mechanosensation in the peripheral
nervous system and in neurotransmission in the central
nervous system. Here, we briefly review advances in our
understanding of ASICs, their potential contributions to
disease, and the possibility for their therapeutic
modification.
Introduction
Acid-sensing ion channels (ASICs) represent an H+-gated
subgroup of the degenerin/epithelial Na+ channel (DEG/
ENaC) family of cation channels (Box 1, Figure 1). In this
article we review emerging roles of ASICs, both physiological and pathological, and discuss their therapeutic potential. The sections are organized as questions to emphasize
what we consider to be important areas for future research.
Where are ASICs located?
Understanding the physiological function of ASICs
requires knowledge of their cellular and subcellular location. In the peripheral nervous system (PNS), several
recent studies show that some ASIC subunits are well
positioned to function as sensory receptors, for example
in specialized nerve endings (Box 2, Figure 2a). Reports on
their location in the brain have differed: some evidence
places ASIC1a at synapses [1,2], most likely in the postsynaptic membrane, whereas other data suggest it might
be located throughout the neuron [3,4] (Box 2, Figure 2b).
These observations raise the possibility that ASICs function at multiple subcellular locations. More work is needed
to learn how the distribution and proportions of all subunits vary in different cells and under different physiological conditions.
Corresponding authors: Wemmie, J.A. (john-wemmie@uiowa.edu);
Welsh, M.J. (michael-welsh@uiowa.edu).
Available online 7 August 2006.
www.sciencedirect.com

What activates ASICs in vivo?


Extracellular protons are crucial for ASIC activation.
ASICs are gated by extracellular protons and are largely
responsible for the acid-evoked currents first observed
many years ago in neurons cultured from the PNS and
CNS [57]. In medium-diameter dorsal root ganglion
(DRG) neurons, acid-activated currents closely match
those from heterologous cells expressing ASIC1, ASIC2
and ASIC3 subunits [8]. Moreover, in such DRG neurons
from mice missing individual ASIC subunits, the H+-gated
currents matched those obtained in Chinese hamster ovary
(CHO) cells expressing the subunits that remained [810]
(Figure 1b). The situation is similar in CNS neurons, where
H+-gated currents probably represent a mix of complexes
containing ASIC1a, ASIC2a and ASIC2b [7,11,12]. However, in contrast to the PNS, in CNS neurons ASIC1a has a
particularly important role. Disrupting the gene encoding
ASIC1a eliminated pH-5-evoked current in cultured hippocampal, amygdala and cortical neurons [2,11]
(Figure 1b). And an antagonist specific for ASIC1a homomultimers (PcTX-1, from tarantula venom) inhibited a
significant proportion of the acid-evoked current in cultured hippocampal neurons [13,14].
Thus, in vivo ASICs are likely to affect neuron function
in response to extracellular acid. But this proposition
raises two questions. First, how sensitive are these channels to pH reductions? Second, does in vivo pH fall to levels
sufficient for ASIC activation? In cultured neurons, pH
values 7.2 [10,15,16] can activate these currents, suggesting that they can respond to very small pH changes. In the
PNS, extracellular pH is well known to drop below 7 under
various conditions [1722]. In the CNS, electrical stimulation can lower extracellular pH in hippocampal slices and
in vivo [23,24]. Synaptic vesicle pH is 5.25.7 [25], and so
during neurotransmission synaptic cleft pH can fall. In
retinal ganglion synapses, neurotransmission lowers pH
sufficiently to inhibit presynaptic Ca2+ channels, suggesting that pH changes in the synaptic cleft can be significant
[26,27]. In the postsynaptic membrane, ASICs are ideally
positioned to respond to these localized pH reductions
(Figures 2b and 3).

0166-2236/$ see front matter 2006 Elsevier Ltd. All rights reserved. doi:10.1016/j.tins.2006.06.014

Review

TRENDS in Neurosciences

Box 1. Acid-sensing ion channels


ASICs are members of the DEG/ENaC family, which is characterized
by two transmembrane domains, a large cysteine-rich extracellular
domain, and several small conserved amino acid motifs (Figure 1a;
for recent review see Refs [54,55,83]). At least three genes (Accn2,
Accn1 and Accn3) and their alternatively spliced transcripts (ASIC1a,
ASIC1b, ASIC2a, ASIC2b and ASIC3) produce subunits that are
activated by acid or modulate other ASIC subunits [84] (Figure 1b).
They are preferentially permeable to Na+, but to a lesser extent can
also conduct other cations (e.g. Ca2+, K+, Li+ and H+). ASICs are
expressed in neurons and assemble into homomultimeric and
heteromultimeric complexes. It is not known precisely how many
subunits are required to form a channel, but stoichiometric studies
with other DEG/ENaC channels suggest four to nine [85,86]. In
heterologous cells, ASIC homomultimeric channels vary in terms of
activation and desensitization kinetics (Figure 1b) and pH sensitivity.
For example, in one study, the pH values required for half-maximal
activation (pH0.5) of mouse ASIC subunits were 6.8 (ASIC1a), 6.2
(ASIC1b), 4.9 (ASIC2a) and 6.6 (ASIC3) [8]. There might also be
differences in pH sensitivity between species, for example ASIC3
might be the most sensitive subunit in rat (pH0.5 = 6.7) [10].
Coexpressing various subunits produces heteromultimers, with
H+-gated currents revealing characteristics that in some cases reflect
average properties of contributing subunits and in other cases
properties that lie outside the range of the individual subunits (e.g.
rate of desensitization) [8,9,50,87]. Subunit composition also alters
ion selectivity; for example ASIC1a homomultimeric complexes are
permeable to Ca2+, whereas ASIC2-containing heteromultimeric
complexes are not [65]. Two additional genes, encoding ASIC4 [88]
and BLINaC [89], are expressed in neurons but have not yet been
shown to produce or modulate H+-evoked currents. In addition,
ENaC subunits have been detected in some neurons, where they
might contribute to sensory function [90].

Despite these compelling arguments, attempts so far


have failed to detect ASIC activation during neurotransmission in hippocampal slices or in cultured neurons [2,3].
Our interpretation of these studies is either that ASIC1a is
not activated during neurotransmission or that appropriate tests of ASIC activation at the synapse have not been
identified. More information is needed to understand
clearly the role of ASIC1a in synaptic physiology.
What else does the large extracellular ASIC domain do?
Are there other ligands?
The large size of the ASIC extracellular domain (318 of
the 528 residues in ASIC1a), with its 14 conserved cysteine
residues, suggests a rigid structure and evokes the image
of an antenna protruding above the channel pore
(Figure 1a). It would be surprising if such a complex
structure exists solely to sense protons, so we speculate
that other ligands might exist. Perhaps unrecognized
ligands are released during neurotransmission. Fuelling
this speculation, FMRFamide, a neurotransmitter peptide
from molluscs, modulates activation of ASICs by protons
[28]. Many other factors also interact with the extracellular
domain to modulate H+-dependent channel activity; these
include Zn2+ [29], Ca2+ [16], lactate [30], proteases [31],
arachidonic acid [32], lead [33] and redox reagents [34]
(Figure 3). Conversely, why do protons fail to activate some
ASICs (ASIC2b and ASIC4)? Perhaps these subunits contribute only to heteromultimeric channels that also contain
H+-gated subunits, or maybe they have specific ligands of
their own. Certainly there are other DEG/ENaC subunits
for which we do not yet know the ligands for example, the
related subunits from fish, some of which are insensitive to
www.sciencedirect.com

Vol.29 No.10

579

acid [35,36], and the Drosophila [37] and Caenorhabditis


elegans DEG/ENaC subunits.
We speculate that ASIC extracellular domains will turn
out to possess additional functions. For example, they
might bind components of the extracellular matrix.
Genetic data from C. elegans suggest that interactions
with the extracellular matrix are functionally important
for mechanosensitive DEG/ENaC channels (e.g. Mec-4 and
Mec-10) [38]. Because ASICs are implicated in mechanosensation, binding to the extracellular matrix could support a tethered mechanism by which movement regulates
channel opening. In addition, if the extracellular domain
protrudes a substantial distance from the cell surface, it
might contact proteins on other cells, an interaction that
could influence function.
What is the physiological significance of ASICs in the
PNS?
Their location and properties make ASIC subunits attractive candidates to serve as H+-gated nociceptors. It is
known that peripheral pH falls to <7 with inflammation,
infection, ischemia, hematomas and exercise [1722].
Moreover, such acidosis is well recognized to activate
nociceptors and produce pain in humans that can be attenuated by the DEG/ENaC inhibitor amiloride [3941].
Variations in subunit composition would also enable
ASICs to detect a broad pH range and thereby sense
variations in the intensity of noxious stimuli [8]. In addition, inflammatory mediators such as nerve growth factor
(NGF), 5-hydroxytryptamine (5-HT or serotonin), interleukin-1, bradykinin and brain-derived neurotrophic factor
(BDNF) can stimulate ASIC transcription, which perhaps
contributes to the pain-enhancing effects of these mediators [4244].
Although it seemed that studying gene-targeted mice
would clarify the role of ASICs in acid-induced nociception,
reports to date give a mixed picture. Some data suggest
that loss of ASIC3 reduces the painful response to an acidic
stimulus. For example, in the isolated skinnerve preparation a pH 5 stimulus generated a reduced response in Cmechanoheat fibers of ASIC3 null animals [45]. In addition, repeated injection of acidic saline into muscle produced long-lasting mechanical hyperalgesia in wild-type
but not ASIC3/ mice [46]. Amiloride pretreatment also
attenuated the response in wild-type animals. Based on
these data, we expected that disrupting ASIC3 would blunt
the behavioral response to acid injection into the paw.
However, ASIC3 null mice behaved like wild-type mice
[45]. Even more puzzling, in a different ASIC3 knockout
mouse, intraperitoneal acid injection elicited a greater
behavioral response than in wild-type mice [47]; painful
mechanical and thermal stimuli also generated an
enhanced behavioral response. Although ASIC1a is as
sensitive to pH as ASIC3 and is expressed in DRG neurons,
ASIC1a null mice were no different from wild-type animals
either in the skinnerve preparation or in tests of mechanical hyperalgesia [46,48]. In transgenic mice overexpressing a dominant-negative ASIC3, the response to thermal
stimuli was normal, but the animals showed an increased
response to intraperitoneal acid injection and mechanical
hypersensitivity after inflammatory stimuli [49].

580

Review

TRENDS in Neurosciences Vol.29 No.10

Figure 1. ASIC subunits and activation by extracellular acid. (a) Topology of ASIC1a protein. Adapted, with permission, from Ref. [54]. (b) Acid-evoked currents in
heterologous cells (Cos-7 cells) expressing the indicated ASIC subunits. (c) H+-evoked currents in cultured wild-type dorsal root ganglion (WT DRG) neurons (i) resemble
currents from heterologous cells expressing a combination of ASIC subunits (ii). Currents from DRG neurons cultured from ASIC3/ mice (iii) resemble currents produced
by coexpressing ASIC1a and ASIC2a in Cos-7 cells (iv). (d) Loss of ASIC1a eliminates pH-5-evoked current in neurons cultured from the CNS (hippocampus), whereas loss of
ASIC2 has more subtle effects. Panels (bd) adapted, with permission, from Ref. [2], Ref. [8] (2002) National Academy of Sciences, and Ref. [11].

At present it seems safe to say that ASICs can modulate


the response to some painful stimuli. But it remains difficult to make simple generalizations about the contribution
specific ASICs make to acid-induced nociception or other
painful stimuli, and we are left with several questions.
First, do ASICs serve as receptors for acid that elicits
peripheral pain? This seems likely in some cases, but
the aforementioned data leave much uncertainty. Contributions from multiple different ASIC subunits, the assays
utilized, ASIC-associated proteins, compensatory changes
and mouse strain differences could all cloud
interpretations.
Second, can ASICs produce a response to sustained
acidosis? ASIC currents are largely transient, whereas
nociceptive acidosis can persist for hours to days. Because
they rapidly desensitize, some have considered ASICs best
suited to respond to rapid, transient pH fluctuations.
However, ASICs can manifest a persistent current
[8,50], particularly when activated in the presence of neuropeptides such as NPFF or FMRFamide [28]. In addition,
an overlap between activation and desensitization kinetics
suggests that even mild pH changes lead to a small but
persistent ASIC-mediated depolarization [4,15,51]. Thus,
it seems likely that chronic acidic conditions foster ASIC
activity.
Third, why are acid-activated currents not detected in
peripheral endings of mechanoreceptive nerve fibers in
skin? In culture, large-diameter mechanosensory DRG
neurons express ASICs and manifest H+-gated currents.
www.sciencedirect.com

Moreover, in the skin, peripheral nerve endings contain


ASICs, which are required for normal mechanoreceptor
function [45,52]. Yet paradoxically, acid does not elicit
activity from these endings in a skinnerve preparation
[53]. Possibly ASICs at these sites are H+ gated under some
conditions, but the resulting current fails to depolarize
membranes sufficiently for action potentials to occur.
Alternatively, in mechanosensory nerve terminals, extracellular and intracellular proteins perhaps tether ASICs to
confer touch sensitivity and in so doing mask pH-sensitive
sites or somehow preclude activation by H+. It is also
possible that ASIC activation might require more than
one stimulus at these sites. These considerations suggest
that ASICs might be strongly influenced by the local
environment and perhaps by their interacting partners.
What is the role of ASICs in mechanosensation?
The similarity between ASICs and the mechanoreceptors
(Mec-4 and Mec-10) from C. elegans suggested that ASICs
might contribute to mechanosensation (for more detailed
reviews, see Refs [54,55]). Initially, we anticipated that
disrupting the ASICs might cause a striking loss of function akin to disrupting DEG/ENaC subunits in C. elegans
[56]. But the ASIC effects were more subtle and in some
cases perplexing [54]. In one study, the sensitivity of lowthreshold, rapidly adapting mechanoreceptors was
reduced in ASIC2 null animals [52], but another study
using a different ASIC2 null reported no effect [57].
ASIC1a null mice showed normal responsiveness in a

Review

TRENDS in Neurosciences

Box 2. Localization of ASIC subunits


In the PNS, ASIC subunits localize to neurons innervating skin
[45,52,91], heart [15], gut [92,93] and muscle [94]. They have also
been detected in the eye [95,96], ear [97,98], taste buds [99] and
bone [73]. The localization has been characterized most thoroughly
in cutaneous neurons. ASIC subunits have been detected in large,
medium and small DRG neurons, although the subunit composition
varies among DRG neurons of different sizes, producing heterogeneity in H+-evoked currents [100102]. Immunocytochemical
studies find both ASIC2 and ASIC3 in specialized cutaneous nerve
endings, such as Meissners corpuscles (Figure 2a), palisades of
lanceolate fibers that surround hair shafts, Pilo-Ruffini nerve endings and Merkel cells, and in small free nerve endings in the
epidermis [45,52].
In contrast to the PNS, transcript analysis in the brain shows that
ASIC1a, ASIC2a and ASIC2b have relatively widespread distribution
[61,62,103,104]. ASIC3 is abundantly expressed in DRG and spinal
cord, but little or no ASIC3 is expressed in rodent brain [105,106].
ASIC1a immunostaining and western blotting suggest that ASIC1a
subunits might be particularly enriched in some brain regions,
including the glomerulus of the olfactory bulb, whisker barrel
cortex, cingulate cortex, striatum, nucleus accumbens, amygdala
and cerebellar cortex [1]. At the subcellular level, ASIC1a is present
in the cell body and preferentially localized to dendrites and
dendritic spines (Figure 2b), where it colocalized with the postsynaptic density protein PSD-95. These studies benefited from use
of ASIC1a null mice as a negative control. Consistent with this work,
ASIC1a was also enriched in synaptosome-containing fractions
from brain [13,63]. However, another immunohistochemical study
suggested that ASIC1a might be present in membranes throughout
neurons, including axons, with no preferential distribution to
synapses [3]. Although differences in antibody specificity might
explain the incongruity, additional studies are needed to clarify the
precise location of ASIC1a. Compared with ASIC1a, less is known
about the distribution of ASIC2, although it has been suggested that
ASIC2a is present in the postsynaptic membrane of granule cells
and Purkinje cells of the cerebellum [107]. ASIC1, ASIC2, ASIC3 and
ASIC4 have also been detected in the retina, where pH has an
important role in synaptic physiology [27,95,96,108,109].

skinnerve preparation [48], whereas ASIC3/ mice


manifested increased sensitivity of rapidly adapting
mechanoreceptors [45]. Moreover, mice transgenic for a
dominant-negative ASIC3 that inhibited much of the acidevoked current showed increased behavioral responses to
mechanical stimuli [49]. In splanchnic colonic afferents
and vagal gastroesophageal afferents, disrupting ASIC1a

Vol.29 No.10

581

increased the mechanical sensitivity, disrupting ASIC2


had varied effects depending on the location, and disrupting ASIC3 reduced the mechanosensitivity of all afferent
classes except gastroesophageal mucosal receptors [48,58].
Interestingly, mechanical stimulation of cultured DRG
neurons revealed no difference between ASIC2 and/or
ASIC3 null mutants and wild-type controls [59].
How can these seemingly disparate observations be
reconciled into a unifying hypothesis? The data indicate
that ASICs influence mechanosensation. But their role
seems complex. At present there is no direct evidence
establishing them as mechanosensors. Perhaps redundancy of DEG/ENaC channels and/or the presence of other
sensory receptors including transient receptor potential
(TRP) channels preclude simple all-or-nothing effects. Perhaps ASICs somehow modulate transduction by other
molecules. Perhaps various ASICs can have activating
or inhibiting roles depending on their partners and locations. Perhaps compensatory effects occur in knockout and
transgenic animals. Or perhaps ASIC activation by
mechanical stimuli requires a specific intracellular and/
or extracellular milieu. To understand how ASICs influence mechanical sensation, we need to know whether they
can be gated by mechanical stimuli and how they influence
other sensors.
What is the physiological significance of ASICs in the
CNS?
Based on studies of mice with targeted disruption of the
gene encoding ASIC1a and transgenic mice overexpressing
ASIC1a, we know more about ASIC1a than the other ASIC
subunits. Loss of ASIC1a disrupted hippocampal-dependent LTP and spatial memory in the Morris water maze [2].
These results indicate that ASIC1a modulates synaptic
plasticity and contributes to learning and memory. However, the spatial learning effects were modest [1], perhaps
because hippocampal pyramidal cells express relatively
low levels of ASIC1a [1].
By contrast, the molecular layer of the cerebellar cortex
expresses abundant ASIC1a and acidosis elicits large H+gated currents in Purkinje cells [1,32]. Accordingly, loss of

Figure 2. Subcellular distribution of ASICs. (a) ASIC3 immunostaining in a Meissners corpuscle. S100 immunostaining was used to identify the corpuscle positively. Scale
bar, 100 mm. Adapted, with permission, from Ref. [45]. (b) ASIC1a immunostaining in dendritic spines. In a cultured hippocampal neuron transfected with cDNA constructs
expressing green fluorescent protein (GFP, green) and ASIC1a (red), ASIC1a appears in the spinous processes extending off the dendritic tree, suggesting it is closely
associated with the postsynaptic membrane. Scale bar, 10 mm.
www.sciencedirect.com

582

Review

TRENDS in Neurosciences Vol.29 No.10

Figure 3. Model for ASIC1a activation at the synapse. In the postsynaptic membrane, ASICs might be well positioned to respond to protons released from presynaptic neurotransmitter (NT)-containing vesicles and also from other sources.
Various extracellular modulators and intracellular ASIC-interacting proteins
(boxes) might influence the response, which would be expected to depolarize
the membrane potential and raise intracellular Ca2+ concentration, perhaps
influencing other receptors and signaling proteins. This model makes two predictions that should be tested experimentally. First, it predicts that pH will fall in the
synaptic cleft during neurotransmission. This technically challenging measurement might now be feasible with new pH-sensitive fluorophores. Second, this
model predicts that ASIC1a currents will be activated during neurotransmission.
So far, studies in brain slices [2] and cultured neurons [3] have not detected them.
There are many potential explanations for these negative results, but they hint that
the story is more complex; perhaps ASICs require additional conditions for
activation or interactions with other proteins through the extracellular domain. For
abbreviations, see Table 1 and the main text.

ASIC1a considerably impaired cerebellum-dependent


learning assessed by eye-blink conditioning [2]. ASIC1a
expression and acid-evoked currents are also substantial
in the amygdala complex, which is important in fearrelated behavior [1,6062]. Consistent with this, mice with
a disruption of the gene encoding ASIC1a showed a
reduced fear response (e.g. freezing behavior) to both cued
and context-specific Pavlovian fear conditioning
(Figure 4a) [1]. Conversely, overexpressing ASIC1a in
the amygdala and elsewhere in the brain increased H+evoked currents and enhanced context fear conditioning,
providing an animal model of acquired anxiety [63].
Yet the underlying mechanisms linking ASIC1a function to these behavioral effects remain poorly understood.
Because ASIC1a conducts Ca2+ [62,64,65] and interacts
with Ca2+/calmodulin-dependent kinase (CaMKII) [66], we
wonder whether a resulting increase in Ca2+ is responsible.
Other questions also arise. Does ASIC1a augment neurotransmission? Do ASICs stimulate or inhibit action potentials [4]? Are ASIC1a homomultimers or heteromultimers
responsible for these effects? Do ASIC2 and other DEG/
ENaC subunits modify the function of ASIC1a in brain, or
do they have independent functions?
What is the physiological significance of interactions
with other proteins?
Most membrane proteins reside in macromolecular complexes, and ASICs are no exception: several ASIC-interacting proteins have already been identified (Table 1) and
more are likely to exist. Possibly the interacting partners
dictate the diverse physiological roles of these channels.

Figure 4. Loss of ASICs disrupts conditioned fear and pain. (a) On day 1, animals received aversive footshocks, which were paired with the training environment (context) or
a tone. On day 2, the conditioned freezing responses to the context and tone were measured. Disruption of the gene encoding ASIC1a reduced the freezing response on day
2 to both context and tone. Adapted, with permission, from Ref. [1] (2003) Society for Neuroscience. (b) Paw withdrawal before and after intramuscular injection of acid
(pH 4.0). Disruption of the gene encoding ASIC3 reduced post-injection hyperalgesia. Adapted, with permission, from Ref. [45]. Asterisks indicate P < 0.05.
www.sciencedirect.com

Review

TRENDS in Neurosciences

Vol.29 No.10

583

Table 1. Interactions of ASIC subunits with scaffolding proteins, and effects on ASIC propertiesa
Protein
PICK1
CaMKII
PSD-95
Lin7B
MAGI-1b
PIST
CIPP
Stomatin
Stomatin
Stomatin
NHERF
Annexin II/p11 1

ASIC subunits
1, 2
1
3
3
3
3
3
1
2
3
3
1

Site of interaction b
C terminus
NR
C terminus
C terminus
C terminus
C terminus
C terminus
NR
NR
NR
C terminus
N terminus

Current amplitude c
$
"
#
"
"
"
"
$
$
#
"
"

Gating properties
NR
NR
$
$
NR
NR
# pH sensitivity
$
" desensitization
$
$
$

Surface expression
NR
NR
#
"
NR
NR
NR
NR
NR
$
"
"

Refs
[6769]
[66]
[106]
[106]
[106]
[106]
[110]
[111]
[111]
[111]
[112]
[70]

Symbols: ", increases; #, decrease; $, no change. Abbreviations: CaMKII, Ca2+/calmodulin-dependent kinase II; CIPP, channel-interacting PDZ domain protein; Lin7B,
abnormal cell lineage 7B; MAGI-1b, membrane-associated guanylate kinase with inverted orientation-1b; NHERF, Na+/H+ exchanger regulatory factor 1; NR, not reported;
PICK1, protein interacting with C-kinase 1; PIST, PDZ-domain protein interacting specifically with TC10; PSD-95, postsynaptic density protein 95.
b
Where investigated, binding to the C terminus of ASIC subunits has been through the PDZ domain of the indicated protein.
c
Peak whole-cell current amplitude in response to acid application.

We imagine a dynamic interplay with multiple proteins


shuttling ASICs to specialized sensory receptors and
synapses, or stabilizing them once they are there. They
might regulate ligand binding, channel opening and closing, and second-messenger signaling. These interactions
could also regulate protein turnover, or the assembly of
channel subunits into homomeric and heteromeric complexes. Interactions with other channels and transporters
are also a possibility.
Although current knowledge is inadequate, it does support several conclusions. First, not all ASICs interact with
the same partners. For example, protein interacting with C
kinase (PICK)1 interacts directly with ASIC1 and ASIC2,
but not ASIC3 [6769]. Second, some interactions occur via
the C-terminal PDZ domains (Table 1), at least one (with
annexin) occurs via the N terminus [70], and for others the
interacting sites have not yet been identified. Third, at
least two interactions (with PICK1 and CaMKII) facilitate
channel phosphorylation [69,71]. Fourth, some interactions affect channel gating properties whereas others
might increase or decrease channel density on the cell
surface (Table 1). Each of these interactions might have
important consequences, but much remains to be learned.
For example, which interactions occur in vivo, and what
are their physiological effects? A detailed knowledge of
these interactions is crucial if we are to understand how
ASICs function in complex systems.
What are the roles of ASICs in pathophysiology,
and are there therapeutic opportunities?
From what we have learned so far, there is an emerging
appreciation for how ASICs might contribute to disease
and to novel therapies. Here we consider a few examples.
Pain
There is an enormous need for better medications to treat
both acute and chronic pain, so the possibility of targeting
ASICs has substantial appeal. The idea that inhibiting
ASICs might relieve pain in humans received a significant
boost when the DEG/ENaC antagonist amiloride attenuated the pain of intradermal acid infusion [40,41]. Although
amiloride can have other effects, it failed to inhibit capsaicin-induced pain, and capsaizapine, an inhibitor of vanilloid
receptor 1 (VR1), did not relieve acid-evoked pain. These
www.sciencedirect.com

results suggested that ASICs, rather than VR1, mediated


pain induced by acid injection. In addition, an agent (A317567) that inhibits ASIC1, ASIC2 and ASIC3 subunits
also attenuated post-operative skin-incision pain in rats
[72]. Furthermore, non-steroidal anti-inflammatory drugs
(NSAIDs) reduced ASIC expression and inhibited ASIC
currents in rat sensory neurons [42,43].
One of the first suggestions that ASICs might be
involved in visceral pain came with the observation that
cardiac afferents express ASIC3 currents, which might
mediate the pain of ischemic angina [10,15]. In addition,
recent data show that ASICs affect the response to gut
distension [48,58]. Because mechanical stimulation can
influence autonomic control of digestion and can evoke
feelings of bloating, cramping and pain, ASICs might be
good targets for combating visceral discomfort. ASICs
might also be involved in pain at several locations. Their
expression in nerves innervating joints and bone positions
them to sense the acidosis that arises in chronic inflammatory arthritides and cancer metastases [73,74]. ASICs
in muscle afferents might detect the lactic acidosis of
severe exercise. Sickle cell disease can cause ischemia with
crippling pain; perhaps ASICs detect the resulting pH
reduction [75].
We wonder whether targeting either specific ASIC subunits or all ASICs might yield novel treatments for common pain syndromes. Attempts to develop toxins [13,76]
and small molecule ASIC inhibitors [72] could yield new
tools to investigate these possibilities. However, as we
have already described, although several studies suggest
ASIC inhibition might reduce pain, some mouse data
suggest disrupting ASICs might increase sensitivity to
painful stimuli. The explanations for such differences
remain unclear, but might include variability in genetic
background or species, differences in testing paradigms,
and environmental variability. Or it might be that ASICs
affect pain in ways more complex than originally appreciated. We clearly need to know more about the physiological role of ASICs in pain before we can take full
advantage of their therapeutic potential.
Ischemic stroke
Recent studies have generated significant interest in
ASICs as possible therapeutic targets for stroke. Severe

584

Review

TRENDS in Neurosciences Vol.29 No.10

cerebral ischemia can cause pH to drop to 6.3 and below


[77,78]. Because Ca2+ overload causes toxicity in the
ischemic brain, and because ASIC1a homomultimers conduct Ca2+ [65], ASIC1a activation could contribute to the
cell damage and death of cerebral ischemia. Enhancement
of ASIC1a function by activation of NMDA receptors, lack
of oxygen or glucose, or accumulation of lactate or arachidonic acid might further enhance ASIC1a-mediated neurotoxicity [66]. The hypothesis that ASIC1a contributes to
ischemic toxicity was recently tested in cells heterologously
expressing ASIC1a and in hippocampal neurons [6466].
Neurons lacking ASIC1a and cells treated with amiloride
or PcTX-1 resisted acidosis-induced injury. Moreover,
PcTX-1 diminished the effects of NMDA-induced cell death
[64,66]. Another key finding came from a mouse model in
which middle cerebral artery occlusion causes an ischemic
stroke [64]. Remarkably, disrupting ASIC1a reduced
infarct volume by 60%. Because pH remains low for hours
after a stroke, ASIC1a inhibitors might be beneficial for
some time after an ischemic event [79]. These results
recommend ASIC1a inhibitors as a potential new treatment for stroke. Likewise, we wonder whether increasing
ASIC2a expression might be neuroprotective by forming
heteromultimers of ASIC1a and ASIC2a that do not conduct Ca2+. Consistent with this possibility, ASIC2a expression is increased following global ischemia, perhaps
serving a neuroprotective function [80].
Psychiatric disease
We currently know of no psychiatric disease caused by a
loss or gain of ASIC function. But several factors lead us to
predict that a link to psychiatric disease will be discovered.
First, ASIC1a is located at synapses and modulation of
synaptic physiology is likely to have behavioral consequences. Second, expression of ASIC1a, ASIC2a and
ASIC2b in the brain is robust. Third, the loss of ASIC1a
in mice had several behavioral effects, suggesting that
similar consequences might carry over to humans.
What psychiatric manifestations could be associated
with ASIC1a mutations? ASIC1a is particularly abundant
in the amygdala and other brain regions implicated in
anxiety, and the loss of ASIC1a has pronounced effects
on fear-related behavior [1]. Moreover, because the anatomy and physiology of fear conditioning are well conserved
between species and ASIC1a is expressed in the human
amygdala [1,63], we anticipate that disrupted ASIC1a
function would reduce fear in humans. Because reduced
fear has been suggested to increase risk-taking behavior
[81]; this personality trait might be associated with a loss of
ASIC1a function. Conversely, analysis of mice overexpressing ASIC1a suggests that humans with gain-of-function
mutations might be prone to fearful behavior. They might
also be supersensitive to acidosis. Interestingly many
patients with panic disorder have panic attacks when
breathing CO2 [82]. Because CO2 lowers brain pH, we
wonder whether ASIC1a might have a role in CO2-evoked
panic.
Irrespective of whether ASIC genes contribute to disease pathogenesis, manipulating ASIC1a function pharmacologically might provide a novel strategy for reducing
fear and anxiety in patients. The relative abundance of
www.sciencedirect.com

ASIC1a in the fear circuit and its robust effect on


fear-related behavior suggest that ASIC1a antagonists
might offer advantages over current treatments, which
target the GABA and 5-HT systems. In mice, the effect
of ASIC1a on hippocampus-dependent memory was relatively subtle; thus, in humans the effect of ASIC1a antagonists on declarative memory might be small [2].
Furthermore, several other non-fear-related behaviors
were intact in the ASIC1a null mice [1,2], suggesting
ASIC1a antagonists might reduce anxiety with relative
specificity.
In conclusion, our knowledge of ASICs is rapidly
increasing. We have learned much about their localization,
and their biochemical, physiological and pathophysiological effects. This knowledge raises additional mechanistic
questions. However, it also increases our appreciation that
ASICs might be excellent targets for treating neurological
disease.
Acknowledgements
J.A.W. was supported by the Department of Veterans Affairs Advanced
Research Career Development Award, NARSAD Young Investigatory
Award, and ADAA Young Investigator Award. M.J.W. is an Investigator
of the HHMI.

References
1 Wemmie, J.A. et al. (2003) Acid-sensing ion channel 1 is localized in
brain regions with high synaptic density and contributes to fear
conditioning. J. Neurosci. 23, 54965502
2 Wemmie, J.A. et al. (2002) The acid-activated ion channel ASIC
contributes to synaptic plasticity, learning, and memory. Neuron
34, 463477
3 Alvarez de la Rosa, D. et al. (2003) Distribution, subcellular
localization and ontogeny of ASIC1 in the mammalian central
nervous system. J. Physiol. 546, 7787
4 Vukicevic, M. and Kellenberger, S. (2004) Modulatory effects of acidsensing ion channels (ASICs) on action potential generation in
hippocampal neurons. Am. J. Physiol. Cell. Physiol. 287, C682C690
5 Gruol, D.L. et al. (1980) Hydrogen ions have multiple effects on the
excitability of cultured mammalian neurons. Brain Res. 183, 247252
6 Krishtal, O.A. and Pidoplichko, V.I. (1981) A receptor for protons in
the membrane of sensory neurons may participate in nociception.
Neuroscience 6, 25992601
7 Varming, T. (1999) Proton-gated ion channels in cultured mouse
cortical neurons. Neuropharmacology 38, 18751881
8 Benson, C.J. et al. (2002) Heteromultimerics of DEG/ENaC subunits
form H+-gated channels in mouse sensory neurons. Proc. Natl. Acad.
Sci. U. S. A. 99, 23382343
9 Xie, J. et al. (2002) DRASIC Contributes to pH-gated currents in large
dorsal root ganglion sensory neurons by forming heteromultimeric
channels. J. Neurophysiol. 87, 28352843
10 Sutherland, S.P. et al. (2001) Acid-sensing ion channel 3 matches the
acid-gated current in cardiac ischemia-sensing neurons. Proc. Natl.
Acad. Sci. U. S. A. 98, 711716
11 Askwith, C.C. et al. (2004) ASIC2 modulates ASIC1 H+-activated
currents in hippocampal neurons. J. Biol. Chem. 279, 1829618305
12 Baron, A. et al. (2002) ASIC-like, proton-activated currents in rat
hippocampal neurons. J. Physiol. 539, 485494
13 Escoubas, P. et al. (2000) Isolation of a tarantula toxin specific for a
class of proton-gated Na+ channels. J. Biol. Chem. 275, 2511625121
14 Chen, X. et al. (2005) The tarantula toxin psalmotoxin 1 inhibits acidsensing ion channel (ASIC) 1a by increasing its apparent H+ affinity.
J. Gen. Physiol. 126, 7179
15 Benson, C.J. et al. (1999) Acid-evoked currents in cardiac sensory
neurons: a possible mediator of myocardial ischemic sensation. Circ.
Res. 84, 921928
16 Immke, D.C. and McCleskey, E.W. (2003) Protons open acid-sensing
ion channels by catalyzing relief of Ca2+ blockade. Neuron 37, 7584
17 Revici, E. et al. (1949) The painful focus. II. The relation of pain to local
physico-chemical changes. Bull Inst. Appl. Biol. 1, 21

Review

TRENDS in Neurosciences

18 Pan, J.W. et al. (1988) Intracellular pH in human skeletal muscle by


1
H NMR. Proc. Natl. Acad. Sci. U. S. A. 85, 78367839
19 Hood, V.L. et al. (1988) Effect of systemic pH on pHi and lactic acid
generation in exhaustive forearm exercise. Am. J. Physiol. 255, F479
F485
20 Issberner, U. et al. (1996) Pain due to tissue acidosis: a mechanism for
inflammatory and ischemic myalgia? Neurosci. Lett. 208, 191194
21 Cobbe, S.M. and Poole-Wilson, P.A. (1980) The time of onset and
severity of acidosis in myocardial ischaemia. J. Mol. Cell. Cardiol. 12,
745760
22 Jacobus, W.E. et al. (1977) Phosphorus nuclear magnetic resonance of
perfused working rat hearts. Nature 265, 756758
23 Krishtal, O.A. et al. (1987) Rapid extracellular pH transients related
to synaptic transmission in rat hippocampal slices. Brain Res. 436,
352356
24 Chesler, M. and Kaila, K. (1992) Modulation of pH by neuronal
activity. Trends Neurosci. 15, 396402
25 Miesenbock, G. et al. (1998) Visualizing secretion and synaptic
transmission with pH-sensitive green fluorescent proteins. Nature
394, 192195
26 DeVries, S.H. (2001) Exocytosed protons feedback to suppress the
Ca2+ current in mammalian cone photoreceptors. Neuron 32, 1107
1117
27 Vessey, J.P. et al. (2005) Proton-mediated feedback inhibition of
presynaptic calcium channels at the cone photoreceptor synapse. J.
Neurosci. 25, 41084117
28 Askwith, C.C. et al. (2000) Neuropeptide FF and FMRFamide
potentiate acid-evoked currents from sensory neurons and protongated DEG/ENaC channels. Neuron 26, 133141
29 Baron, A. et al. (2001) Zn2+ and H+, coactivators of acid sensing ion
channels (ASIC). J. Biol. Chem. 276, 3536135367
30 Immke, D.C. and McCleskey, E.W. (2001) Lactate enhances the acidsensing Na+ channel on ischemia-sensing neurons. Nat. Neurosci. 4,
869870
31 Poirot, O. et al. (2004) Selective regulation of acid-sensing ion channel
1 by serine proteases. J. Biol. Chem. 279, 3844838457
32 Allen, N.J. and Attwell, D. (2002) Modulation of ASIC channels in rat
cerebellar Purkinje neurons by ischaemia-related signals. J. Physiol.
543, 521529
33 Wang, W. et al. (2006) Calcium-permeable acid-sensing ion channel is
a molecular target of the neurotoxic metal ion lead. J. Biol. Chem. 281,
24972505
34 Andrey, F. et al. (2005) Acid sensing ionic channels: modulation by
redox reagents. Biochim. Biophys. Acta 1745, 16
35 Paukert, M. et al. (2004) A family of acid-sensing ion channels from the
zebrafish: widespread expression in the central nervous system
suggests a conserved role in neuronal communication. J. Biol.
Chem. 279, 1878318791
36 Coric, T. et al. (2005) Proton sensitivity of ASIC1 appeared with the
rise of fishes by changes of residues in the region that follows TM1 in
the ectodomain of the channel. J. Physiol. 568, 725735
37 Adams, C.M. et al. (1998) Ripped pocket and pickpocket, novel
Drosophila DEG/ENaC subunits expressed in early development
and in mechanosensory neurons. J. Cell Biol. 140, 143152
38 Gu, G. et al. (1996) Genetic interactions affecting touch sensitivity in
Caenorhabditis elegans. Proc. Natl. Acad. Sci. U. S. A. 93, 65776582
39 Steen, K.H. and Reeh, P.W. (1993) Sustained graded pain and
hyperalgesia from harmless experimental tissue acidosis in human
skin. Neurosci. Lett. 154, 113116
40 Ugawa, S. et al. (2002) Amiloride-blockable acid-sensing ion channels
are leading acid sensors expressed in human nociceptors. J. Clin.
Invest. 110, 11851190
41 Jones, N.G. et al. (2004) Acid-induced pain and its modulation in
humans. J. Neurosci. 24, 1097410979
42 Mamet, J. et al. (2002) Proinflammatory mediators, stimulators of
sensory neuron excitability via the expression of acid-sensing ion
channels. J. Neurosci. 22, 1066210670
43 Voilley, N. et al. (2001) Nonsteroid anti-inflammatory drugs inhibit
both the activity and the inflammation-induced expression of acidsensing ion channels in nociceptors. J. Neurosci. 21, 80268033
44 McIlwrath, S.L. et al. (2005) The sensory mechanotransduction ion
channel ASIC2 (acid sensitive ion channel 2) is regulated by
neurotrophin availability. Neuroscience 131, 499511
www.sciencedirect.com

Vol.29 No.10

585

45 Price, M.P. et al. (2001) The DRASIC cation channel contributes to the
detection of cutaneous touch and acid stimuli in mice. Neuron 32,
10711083
46 Sluka, K.A. et al. (2003) Chronic hyperalgesia induced by repeated
acid injections in muscle is abolished by the loss of ASIC3, but not
ASIC1. Pain 106, 229239
47 Chen, C.C. et al. (2002) A role for ASIC3 in the modulation of highintensity pain stimuli. Proc. Natl. Acad. Sci. U. S. A. 99, 89928997
48 Page, A.J. et al. (2004) The ion channel ASIC1 contributes to visceral
but not cutaneous mechanoreceptor function. Gastroenterology 127,
17391747
49 Mogil, J.S. et al. (2005) Transgenic expression of a dominant-negative
ASIC3 subunit leads to increased sensitivity to mechanical and
inflammatory stimuli. J. Neurosci. 25, 98939901
50 Lingueglia, E. et al. (1997) A modulatory subunit of acid sensing ion
channels in brain and dorsal root ganglion cells. J. Biol. Chem. 272,
2977829783
51 Yagi, J. et al. (2004) Small changes in pH cause sustained current in
acid-sensing ion channel 3, ASIC3. In 2004 Abstract Viewer and
Itinerary Planner, program number 859.3, Society for
Neuroscience, online (http://sfn.scholarone.com/)
52 Price, M.P. et al. (2000) The mammalian sodium channel BNC1 is
required for normal touch sensation. Nature 407, 10071011
53 Lewin, G.R. and Stucky, C.L. (2000) Sensory neuron
mechanotransduction: its regulation and underlying molecular
mechanisms. In Molecular Basis of Pain Induction (Wood, J.N.,
ed.), pp. 129149, Wiley
54 Welsh, M.J. et al. (2002) Biochemical basis of touch perception:
mechanosensory function of degenerin/epithelial Na+ channels. J.
Biol. Chem. 277, 23692372
55 Krishtal, O. (2003) The ASICs: signaling molecules? Modulators?.
Trends Neurosci. 26, 477483
56 Driscoll, M. and Tavernarakis, N. (2000) Closing in on a mammalian
touch receptor. Nat. Neurosci. 3, 12321234
57 Roza, C. et al. (2004) Knockout of the ASIC2 channel in mice does not
impair cutaneous mechanosensation, visceral mechanonociception
and hearing. J. Physiol. 558, 659669
58 Page, A.J. et al. (2005) Different contributions of ASIC channels 1a, 2,
and 3 in gastrointestinal mechanosensory function. Gut 54, 1408
1415
59 Drew, L.J. et al. (2004) Acid-sensing ion channels ASIC2 and ASIC3
do not contribute to mechanically activated currents in mammalian
sensory neurones. J. Physiol. 556, 691710
60 Olson, T.H. et al. (1998) An acid sensing ion channel (ASIC) localizes
to small primary afferent neurons in rats. NeuroReport 9, 11091113
61 Garca-Anoveros, J. et al. (1997) BNaC1 and BNaC2 constitute a new
family of human neuronal sodium channels related to degenerins and
epithelial sodium channels. Proc. Natl. Acad. Sci. U. S. A. 94, 1459
1464
62 Waldmann, R. et al. (1997) A proton-gated cation channel involved in
acid-sensing. Nature 386, 173177
63 Wemmie, J. et al. (2004) Overexpression of acid-sensing ion channel
1a in transgenic mice increases fear-related behavior. Proc. Natl.
Acad. Sci. U. S. A. 101, 36213626
64 Xiong, Z.G. et al. (2004) Neuroprotection in ischemia: blocking
calcium-permeable acid-sensing ion channels. Cell 118, 687698
65 Yermolaieva, O. et al. (2004) Extracellular acidosis increases neuronal
cell calcium by activating acid-sensing ion channel 1a. Proc. Natl.
Acad. Sci. U. S. A. 101, 67526757
66 Gao, J. et al. (2005) Coupling between NMDA receptor and acidsensing ion channel contributes to ischemic neuronal death.
Neuron 48, 635646
67 Hruska-Hageman, A.M. et al. (2002) Interaction of the synaptic
protein PICK1 (protein interacting with C kinase 1) with the nonvoltage gated sodium channels BNC1 (brain Na+ channel 1) and ASIC
(acid-sensing ion channel). Biochem. J. 361, 443450
68 Duggan, A. et al. (2002) The PDZ domain protein PICK1 and the
sodium channel BNaC1 interact and localize at mechanosensory
terminals of dorsal root ganglion neurons and dendrites of central
neurons. J. Biol. Chem. 277, 52035208
69 Baron, A. et al. (2002) Protein kinase C stimulates the acid-sensing ion
channel ASIC2a via the PDZ domain-containing protein PICK1. J.
Biol. Chem. 277, 5046350468

Review

586

TRENDS in Neurosciences Vol.29 No.10

70 Donier, E. et al. (2005) Annexin II light chain p11 promotes functional


expression of acid-sensing ion channel ASIC1a. J. Biol. Chem. 280,
3866638672
71 Leonard, A.S. et al. (2003) cAMP-dependent protein kinase
phosphorylation of the acid-sensing ion channel-1 regulates its
binding to the protein interacting with C-kinase-1. Proc. Natl.
Acad. Sci. U. S. A. 100, 20292034
72 Dube, G.R. et al. (2005) Electrophysiological and in vivo
characterization of A-317567, a novel blocker of acid sensing ion
channels. Pain 117, 8896
73 Jahr, H. et al. (2005) Identification of acid-sensing ion channels in
bone. Biochem. Biophys. Res. Commun. 337, 349354
74 Heppelmann, B. and McDougall, J.J. (2005) Inhibitory effect of
amiloride and gadolinium on fine afferent nerves in the rat knee:
evidence of mechanogated ion channels in joints. Exp. Brain Res. 167,
114118
75 Dickerhoff, R. and von Ruecker, A. (1995) Pain crises in patients with
sickle cell diseases. Pathogenesis, clinical aspects, therapy. Klin.
Padiatr. 207, 321325
76 Diochot, S. et al. (2004) A new sea anemone peptide, APETx2, inhibits
ASIC3, a major acid-sensitive channel in sensory neurons. EMBO J.
23, 15161525
77 Katsura, K. and Siesjo, B.O.K. (1998) Acid-base metabolism in
ischemia. In pH and Brain Function (Kaila, K. and Ransom, B.R.,
eds), p. 563, Wiley-Liss, Inc
78 Siesjo, B.K. et al. (1996) Acidosis-related damage. Adv. Neurol. 71,
209233
79 Pignataro, G. et al. (2005) Neuroprotective time window of ASIC1a
blockade in mouse model of focal cerebral ischemia. In 2005 Abstract
Viewer and Itinerary Planner, program number 669.14, Society for
Neuroscience, online (http://sfn.scholarone.com/)
80 Johnson, M.B. et al. (2001) Global ischemia induces expression of acidsensing ion channel 2a in rat brain. J. Cereb. Blood Flow Metab. 21,
734740
81 Poulton, R. and Menzies, R.G. (2002) Non-associative fear acquisition:
a review of the evidence from retrospective and longitudinal research.
Behav. Res. Ther. 40, 127149
82 Klein, D.F. (1993) False suffocation alarms, spontaneous panics, and
related conditions. An integrative hypothesis. Arch. Gen. Psychiatry
50, 306317
83 Kellenberger, S. and Schild, L. (2002) Epithelial sodium channel/
degenerin family of ion channels: a variety of functions for a
shared structure. Physiol. Rev. 82, 735767
84 Waldmann, R. and Lazdunski, M. (1998) H+-gated cation channels:
neuronal acid sensors in the NaC/DEG family of ion channels. Curr.
Opin. Neurobiol. 8, 418424
85 Firsov, D. et al. (1998) The heterotetrameric architecture of the
epithelial sodium channel (ENaC). EMBO J. 17, 344352
86 Eskandari, S. et al. (1999) Number of subunits comprising the
epithelial sodium channel. J. Biol. Chem. 274, 2728127286
87 Bassilana, F. et al. (1997) The acid-sensitive ionic channel subunit
ASIC and the mammalian degenerin MDEG form a heteromultimeric
H+-gated Na+ channel with novel properties. J. Biol. Chem. 272,
2881928822
88 Akopian, A.N. et al. (2000) A new member of the acid-sensing ion
channel family. NeuroReport 11, 22172222
89 Sakai, H. et al. (1999) Cloning and functional expression of a novel
degenerin-like Na+ channel gene in mammals. J. Physiol. 519, 323
333
90 Drummond, H.A. et al. (1998) A molecular component of the arterial
baroreceptor mechanostransducer. Neuron 21, 14351441

91 Garcia-Anoveros, J. et al. (2001) Transport and localization of the


DEG/ENaC ion channel BNaC1a to peripheral mechanosensory
terminals of dorsal root ganglia neurons. J. Neurosci. 21, 26782686
92 Yiangou, Y. et al. (2001) Increased acid-sensing ion channel ASIC-3 in
inflamed human intestine. Eur. J. Gastroenterol. Hepatol. 13, 891896
93 Schicho, R. et al. (2004) Increased expression of TRPV1 receptor in
dorsal root ganglia by acid insult of the rat gastric mucosa. Eur. J.
Neurosci. 19, 18111818
94 Molliver, D.C. et al. (2005) ASIC3, an acid-sensing ion channel, is
expressed in metaboreceptive sensory neurons. Mol. Pain 1, 35
95 Ettaiche, M. et al. (2004) Acid-sensing ion channel 2 is important for
retinal function and protects against light-induced retinal
degeneration. J. Neurosci. 24, 10051012
96 Ettaiche, M. et al. (2006) Silencing acid-sensing ion channel 1a alters
cone-mediated retinal function. J. Neurosci. 26, 58005809
97 Hildebrand, M.S. et al. (2004) Characterisation of DRASIC in the
mouse inner ear. Hear. Res. 190, 149160
98 Peng, B.G. et al. (2004) Acid-sensing ion channel 2 contributes a major
component to acid-evoked excitatory responses in spiral ganglion
neurons and plays a role in noise susceptibility of mice. J.
Neurosci. 24, 1016710175
99 Ugawa, S. et al. (2003) Amiloride-insensitive currents of the acidsensing ion channel-2a (ASIC2a)/ASIC2b heteromeric sour-taste
receptor channel. J. Neurosci. 23, 36163622
100 Alvarez de la Rosa, D. et al. (2002) Functional implications of the
localization and activity of acid-sensitive channels in rat peripheral
nervous system. Proc. Natl. Acad. Sci. U. S. A. 99, 23262331
101 Ugawa, S. et al. (2005) In situ hybridization evidence for the
coexistence of ASIC and TRPV1 within rat single sensory neurons.
Mol. Brain Res. 136, 125133
102 Ichikawa, H. and Sugimoto, T. (2002) The co-expression of ASIC3 with
calcitonin gene-related peptide and parvalbumin in the rat trigeminal
ganglion. Brain Res. 943, 287291
103 Price, M.P. et al. (1996) Cloning and expression of a novel human
brain Na+ channel. J. Biol. Chem. 271, 78797882
104 Waldmann, R. et al. (1996) The mammalian degenerin MDEG, an
amiloride-sensitive cation channel activated by mutations causing
neurodegeneration in Caenorhabditis elegans. J. Biol. Chem. 271,
1043310436
105 Waldmann, R. et al. (1997) Molecular cloning of a non-inactivating
proton-gated Na+ channel specific for sensory neurons. J. Biol. Chem.
272, 2097520978
106 Hruska-Hageman, A.M. et al. (2004) PSD-95 and Lin-7b interact with
acid-sensing ion channel-3 and have opposite effects on H+-gated
current. J. Biol. Chem. 279, 4696246968
107 Jovov, B. et al. (2003) Immunolocalization of the acid-sensing ion
channel 2a in the rat cerebellum. Histochem. Cell Biol. 119, 437446
108 Brockway, L.M. et al. (2002) Rabbit retinal neurons and glia express a
variety of ENaC/DEG subunits. Am. J. Physiol. Cell Physiol. 283,
C126C134
109 Lilley, S. et al. (2004) The discovery and characterization of a protongated sodium current in rat retinal ganglion cells. J. Neurosci. 24,
10131022
110 Anzai, N. et al. (2002) The multivalent PDZ domain-containing
protein CIPP is a partner of acid- sensing ion channel 3 in sensory
neurons. J. Biol. Chem. 277, 1665516661
111 Price, M.P. et al. (2004) Stomatin modulates gating of acid-sensing ion
channels. J. Biol. Chem. 279, 5388653891
112 Deval, E. et al. (2006) Regulation of sensory neuron-specific acidsensing ion channel 3 by the adaptor protein NHERF-1. J. Biol. Chem.
281, 17961807

Reproduction of material from Elsevier articles


Interested in reproducing part or all of an article published by Elsevier, or one of our article figures?
If so, please contact our Global Rights Department with details of how and where the requested
material will be used. To submit a permission request online, please visit:

www.elsevier.com/locate/permissions
www.sciencedirect.com

Das könnte Ihnen auch gefallen