Top Banner
331 Review www.expert-reviews.com ISSN 1751-2433 © 2010 Expert Reviews Ltd 10.1586/ECP.10.13 Molecular properties Characterization The first molecular identification of acid-sensing ion channels (ASICs) – at that time, known as BNC1, MDEG or BNaC1-2 – was achieved in 1996 and 1997 by Michael Welsh’s, Michel Lazdunski’s and David Corey’s groups [1–3] . However, the essential discovery that these pro- teins form proton-gated channels that could sup- port the proton-activated currents described in the 1980s in neurons [4,5] was made by Michel Lazdunski’s group in 1997 [6] , who called these channels ASICs. They belong to the epithelial Na + channel (ENaC)/DEG/ASIC family of ion channels [7], which comprises ENaC [8,9] involved in taste perception and Na + homeostasis in mam- mals, the neuronal and muscular degenerins of the nematode Caenorhabditis elegans involved in mechanoperception [10–13] and the FMRFamide peptide-gated Na + channel (FaNaC) identified from the invertebrate nervous system [14,15] . The ASIC family in mammals comprises four different genes encoding at least six channel iso- forms (FIGURE 1) . Both ASIC1 and ASIC2 have splice variants, known as ASIC1a (also named ASIC or ASICa or BNaC2a) [3,6] , ASIC1b (also named ASICb or BNaC2b) [16,17] , ASIC2a (also named MDEG1 or BNaC1a or BNC1a) [1–3,18] and ASIC2b (also named MDEG2 or BNaC1b [19] . The ASIC1b isoform has not been described in human so far. One variant of ASIC3 (also named DRASIC or TNaC) [20–23] has been identified in rodents, whereas at least three variants called ASIC3a, b and c (differing in their C-terminal domain) have been reported in humans. ASIC4 (also named SPASIC) [24,25] also has two splice variants in humans, differing only by the deletion of a 19-amino acid region in the extracellular loop. The properties and the physiological relevance of these human splice variants of ASIC3 and ASIC4 remain unclear. Acid-sensing ion channels are assembled in the membrane of neurons as complexes containing accessory proteins, which are important for con- trol of both functional properties and subcellular localization [26–39] . Structure Each ASIC subunit has two hydrophobic trans- membrane regions flanking a large extracellular loop representing more than 50% of the protein, and short intracellular N- and C-termini (FIGURE 1) . Jacques Noël 1 , Miguel Salinas 1 , Anne Baron 1 , Sylvie Diochot 1 , Emmanuel Deval 1 and Eric Lingueglia †1 1 Institut de Pharmacologie Moléculaire et Cellulaire, Institut de NeuroMédecine Moléculaire (IPMC-IN2M) CNRS/ University of Nice-Sophia Antipolis, UMR 6097, 660 Route des Lucioles, Sophia Antipolis, 06560 Valbonne, France Author for correspondence: Tel.: +33 493 953 423 Fax: +33 493 957 708 [email protected] Acid-sensing ion channels (ASICs) form a family of voltage-independent cation channels that predominantly conduct Na + ions, and were identified at the molecular level a little more than a decade ago. ASICs are activated by extracellular acidification within the physiological range, and they form effective proton sensors in both central and peripheral sensory neurons. A combination of genetic and pharmacologic approaches has revealed their implication in an increasing number of physiological and pathophysiological processes – most of them associated with extracellular pH fluctuations, ranging from synaptic plasticity, learning, memory, fear, depression, seizure termination and neuronal degeneration to nociception and mechanosensation. ASICs, therefore, emerge as new potential therapeutic targets in the management of psychiatric disorders, stroke, neurodegenerative diseases and pain. KEYWORDS: acidosis • acid-sensing ion channel • amiloride • brain • chemosensation • dorsal root ganglia • mechanosensation • Na + channels • neurons • pain • protons • synapse Current perspectives on acid‑sensing ion channels: new advances and therapeutic implications Expert Rev. Clin. Pharmacol. 3(3), 331–346 (2010) For reprint orders, please contact [email protected]
16

Current perspectives on acid-sensing ion channels: new advances and therapeutic implications

May 01, 2023

Download

Documents

Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Current perspectives on acid-sensing ion channels: new advances and therapeutic implications

331

Review

www.expert-reviews.com ISSN 1751-2433© 2010 Expert Reviews Ltd10.1586/ECP.10.13

Molecular propertiesCharacterizationThe first molecular identification of acid-sensing ion channels (ASICs) – at that time, known as BNC1, MDEG or BNaC1-2 – was achieved in 1996 and 1997 by Michael Welsh’s, Michel Lazdunski’s and David Corey’s groups [1–3]. However, the essential discovery that these pro-teins form proton-gated channels that could sup-port the proton-activated currents described in the 1980s in neurons [4,5] was made by Michel Lazdunski’s group in 1997 [6], who called these channels ASICs. They belong to the epithelial Na+ channel (ENaC)/DEG/ASIC family of ion channels [7], which comprises ENaC [8,9] involved in taste perception and Na+ homeostasis in mam-mals, the neuronal and muscular degenerins of the nematode Caenorhabditis elegans involved in mechanoperception [10–13] and the FMRFamide peptide-gated Na+ channel (FaNaC) identified from the invertebrate nervous system [14,15].

The ASIC family in mammals comprises four different genes encoding at least six channel iso-forms (Figure 1). Both ASIC1 and ASIC2 have splice variants, known as ASIC1a (also named ASIC or ASICa or BNaC2a) [3,6], ASIC1b

(also named ASICb or BNaC2b) [16,17], ASIC2a (also named MDEG1 or BNaC1a or BNC1a) [1–3,18] and ASIC2b (also named MDEG2 or BNaC1b [19]. The ASIC1b isoform has not been described in human so far. One variant of ASIC3 (also named DRASIC or TNaC) [20–23] has been identified in rodents, whereas at least three variants called ASIC3a, b and c (differing in their C-terminal domain) have been reported in humans. ASIC4 (also named SPASIC) [24,25] also has two splice variants in humans, differing only by the deletion of a 19-amino acid region in the extracellular loop. The properties and the physio logical relevance of these human splice variants of ASIC3 and ASIC4 remain unclear.

Acid-sensing ion channels are assembled in the membrane of neurons as complexes containing accessory proteins, which are important for con-trol of both functional properties and subcellular localization [26–39].

StructureEach ASIC subunit has two hydrophobic trans-membrane regions flanking a large extracellular loop representing more than 50% of the protein, and short intracellular N- and C-termini (Figure 1).

Jacques Noël1, Miguel Salinas1, Anne Baron1, Sylvie Diochot1, Emmanuel Deval1 and Eric Lingueglia†1

1Institut de Pharmacologie Moléculaire et Cellulaire, Institut de NeuroMédecine Moléculaire (IPMC-IN2M) CNRS/ University of Nice-Sophia Antipolis, UMR 6097, 660 Route des Lucioles, Sophia Antipolis, 06560 Valbonne, France†Author for correspondence:Tel.: +33 493 953 423 Fax: +33 493 957 708 [email protected]

Acid-sensing ion channels (ASICs) form a family of voltage-independent cation channels that predominantly conduct Na+ ions, and were identified at the molecular level a little more than a decade ago. ASICs are activated by extracellular acidification within the physiological range, and they form effective proton sensors in both central and peripheral sensory neurons. A combination of genetic and pharmacologic approaches has revealed their implication in an increasing number of physiological and pathophysiological processes – most of them associated with extracellular pH fluctuations, ranging from synaptic plasticity, learning, memory, fear, depression, seizure termination and neuronal degeneration to nociception and mechanosensation. ASICs, therefore, emerge as new potential therapeutic targets in the management of psychiatric disorders, stroke, neurodegenerative diseases and pain.

Keywords: acidosis • acid-sensing ion channel • amiloride • brain • chemosensation • dorsal root ganglia • mechanosensation • Na+ channels • neurons • pain • protons • synapse

Current perspectives on acid‑sensing ion channels: new advances and therapeutic implicationsExpert Rev. Clin. Pharmacol. 3(3), 331–346 (2010)

For reprint orders, please contact [email protected]

Page 2: Current perspectives on acid-sensing ion channels: new advances and therapeutic implications

Expert Rev. Clin. Pharmacol. 3(3), (2010)332

Review Noël, Salinas, Baron, Diochot, Deval & Lingueglia

ASICs are one of the relatively few ion channels with known atomic-scale structure. The tridimensional structure of chicken ASIC1 was solved by Eric Gouaux’s group in 2007 [40,41]. It shows that three subunits need to assemble together to form a channel (Figure 1). The chalice-shaped homotrimer is composed of short N- and C-termini (whose tridimensional structure has not been elucidated so far), two transmembrane helices, a bound Cl- ion and a disulphide-rich extracellular domain. Each subunit has been compared with a clenched hand, which can be further divided into finger, thumb, palm, knuckle, b-ball, wrist in the extracellular domain, and fore-arm as the pore-forming transmembrane domains TM1 and TM2 (Figure 1) [40].

Biophysical properties & gating Selectivity & conductanceAcid-sensing ion channels are low-conductance, depolarizing channels that predominantly conduct Na+ (gNa ~10–15 pS) [6,19,22]. The rapidly desen-sitizing peak current (Figure 1) is always Na+ selective but the additional sustained cur-rent present in ASIC3-containing channels poorly discriminates between Na+ and K+ in human ASIC3 [21], as well as in rat hetero-meric ASIC3+ASIC2b channels [19]. Protons have also been shown to permeate through ASIC1a, which contributes to rundown (tachyphylaxis) of the channel activity upon successive acid stimulations[42]. ASIC1a is, in addition, permeable to Ca2+ [6,17,43]. Activation of ASIC1a increases intracellu-lar free Ca2+concentration in cortical, hip-pocampal and dorsal root ganglia (DRG) neurons, and in cells expressing the recom-binant channel [44–47]. However, direct per-meation of Ca2+ through ASIC1a seems to contribute only modestly to the intracellular Ca2+ rise in neurons [45,47,48], and much of the increase in free Ca2+ could be mediated by indirect secondary mechanisms.

Activation by extracellular pHThe only known activators of ASICs are extracellular protons. However, the ASIC2b and ASIC4 isoforms are not activated by acidic pH on their own but can associate with other ASICs to modu-late their properties and/or their expres-sion [19,49]. The ASIC subunits can assem-ble as either homo- or hetero-trimeric channels, giving rise to a wide range of currents with different sensitivity to exter-nal H+, kinetics and pharmacology, which accounts for the large diversity of native ASIC-like currents observed in neurons [4]. ASIC channels are not voltage-dependent

and their pH0.5

ranges from 4.0 to 6.8, with activation thresh-olds close to pH 7.0 for ASIC1 and ASIC3 (i.e., well in the pathophysio logical pH range) (Table 1). They show steep pH dependence, making them very sensitive sensors of extracellu-lar protons. A rapid drop of extracellular pH triggers a transient inward current that desensitizes rapidly (within seconds) (Figure

1). Interestingly, the reactivation process of the ASIC1a and ASIC2a channels is strongly pH dependent [50]. Pre-existing pH has a profound effect on the ASIC response, and prolonged or slow acidifi cation of the medium results in inactivation of most of the transient response. However, ASIC3-containing channels have an additional slowly activating, sustained component that does not inactivate, so long as the pH remains acidic [20,22]. ASIC3 is actually capable of adapting three different modes of functioning depending on the pH conditions [51]. In the first mode, a rapid decrease in the pH of the extracellular medium

ASIC1aASIC1b

ASIC2aASIC2b

ASIC3

ASIC4

ASIC1a

5s

pH pH

ASIC3

Loop

Loop

AmilorideA-317567 Na +Ca 2+PcTx1

(ASIC1a)

APETx2(ASIC3)

TM

N C

H+H+

H+

H+

H+

H+

TMTM

Out

In

Finger Knuckle

PalmThumb

Wrist

β-ball

A B

C D

Expert Rev. Clin. Pharmacol. © Future Science Group (2010)

Figure 1. Structure and activity of acid-sensing ion channels. (A) The ASIC family is comprised of at least six members that form proton-activated cation channels (except for ASIC2b and ASIC4) with different properties. A rapid drop in extracellular pH triggers a transient inward current that desensitizes within seconds, but ASIC3 has an additional sustained component that does not inactivate while the pH remains acidic. (B) Each ASIC subunit has two hydrophobic transmembrane regions flanking a large extracellular loop, and three subunits assemble together to form a functional channel that predominantly conducts Na+ ions (and some Ca2+ ions for ASIC1a). (C) Amiloride and A-317567 are nonselective (amiloride), nonsubtype-specific blockers of ASICs, while the spider toxin, PcTx1, selectively inhibits ASIC1a channels, and the sea anemone toxin, APETx2, selectively inhibits ASIC3-containing channels. (D) At the atomic level, each subunit looks like a clenched hand [40,41], and it has been proposed that protons binding to the acidic pocket, palm, and wrist domains in the loop open the channel pore through a rotation of the transmembrane helices. ASIC: Acid-sensing ion channels; TM: Transmembrane.

Page 3: Current perspectives on acid-sensing ion channels: new advances and therapeutic implications

www.expert-reviews.com 333

ReviewAcid-sensing ion channels: new advances & therapeutic implicationsTa

ble

1. P

rop

erti

es o

f ac

id-s

ensi

ng

ion

ch

ann

els.

Iso

form

Tiss

ue

dis

trib

uti

on

pH

0.5

acti

vati

on

Inh

ibit

ors

End

og

eno

us

mo

du

lato

rsPh

ysio

log

y Po

ten

tial

th

erap

euti

c im

plic

atio

ns

ASI

C1a

Stro

ng e

xpre

ssio

n in

PN

S, b

rain

, spi

nal c

ord,

re

tina.

Als

o pr

esen

t in

ta

ste

cells

, per

iphe

ral

and

cent

ral a

udito

ry

syst

em, i

mm

une

cells

, lu

ng e

pith

elia

l cel

ls,

bon

e, v

ascu

lar

smo

oth

mus

cle

cells

, glio

mas

6.2

–6

.8A

milo

rid

e an

d d

eriv

ativ

es

ben

zam

il an

d EI

PA [6

],

NSA

IDs

(ibup

rofe

n/

flurb

ipro

fen

) [68

],

A-3

1756

7 [6

7],

nafa

mo

stat

mes

ilate

[70]

, di

aryl

amid

ines

[72]

, Pb2

+

[77]

, Ni2

+ [7

6], Z

n2+ [7

8],

Ca2

+ [6

,63]

, psa

lmot

oxin

-1

(PcT

x1) [

82]

RFa

mid

e p

epti

des

[90]

, d

ynor

phin

s [9

8]

arac

hid

onic

ac

id [5

4,14

2],

lact

ate

[158

],

NO

[143

]

Impo

rtan

t ro

le in

the

CN

S: s

ynap

tic t

rans

mis

sion

and

pl

astic

ity,

lear

ning

and

mem

ory

[101

], in

nate

and

co

nditi

oned

fea

r [9

9,10

6], c

hem

osen

sory

fun

ctio

n of

the

am

ygda

la [1

09],

dep

ress

ion

[110

], c

entr

al m

odul

atio

n of

pa

in t

hrou

gh t

he o

pioi

d sy

stem

[149

], c

entr

al

sens

itiza

tion

to p

ain

in t

he s

pina

l cor

d [5

0,13

2], v

isua

l tr

ansd

uctio

n [1

90].

Als

o in

volv

ed in

prim

ary

hype

ralg

esia

in

mus

cle

[152

], v

isce

ral b

ut n

ot c

utan

eous

m

echa

nose

nsat

ion

[137

,163

], p

ossi

ble

impl

icat

ion

in

chem

otra

nsdu

ctio

n of

low

pH

by

caro

tid b

ody

[184

]

Inhi

biti

on: p

sych

iatr

ic

diso

rder

s (a

nxi

ety,

pan

ic,

dep

ress

ion

), s

tro

ke,

neu

rod

egen

erat

ive

dise

ases

, pai

n A

ctiv

atio

n: s

eizu

re

trea

tmen

t

ASI

C1b

PNS,

tas

te c

ells

, co

chle

ar h

air

cells

, ca

roti

d b

od

y,

imm

une

cells

5.1–

6.2

Am

ilori

de

[16]

, di

aryl

amid

ines

[72]

, Pb

2+ [7

7]

NO

[143

]N

oci

cept

ion?

Pain

?

ASI

C2a

PNS

(incl

udin

g sp

ecia

lized

cut

aneo

us

mec

hano

sens

ory

stru

ctur

es),

bra

in,

spin

al c

ord,

ret

ina,

p

erip

hera

l and

cen

tral

au

dito

ry s

yste

m, t

aste

ce

lls, b

one,

vas

cula

r sm

oot

h m

uscl

e ce

lls,

som

e ex

pres

sion

in

caro

tid

bo

dy,

glio

mas

4.1

–5.0

Am

ilori

de

[18]

, naf

amo

stat

m

esila

te [7

0],

diar

ylam

idin

es [7

2],

Cd2

+ [7

6], A

-317

567

[67]

NO

[143

], Z

n2+ [7

9]M

odu

lati

on o

f A

SIC

1a in

the

CN

S [9

5], v

isua

l tr

ansd

ucti

on [1

88],

cut

aneo

us a

nd v

isce

ral

mec

hano

sens

atio

n [1

28,1

63],

no

ise

susc

eptib

ility

of

hear

ing

[193

], b

aror

ecep

tor

refl

ex [1

83],

may

be

invo

lved

in

sou

r ta

ste

per

cept

ion

[197

,198

], v

ascu

lar

smo

oth

mus

cle

cell

mig

rati

on [2

11]

Con

tro

l of

blo

od

pres

sure

, pai

n?

Not

e g

enet

ic

asso

ciat

ions

wit

h m

ultip

le s

cler

osi

s, a

utis

m

and

men

tal r

etar

dati

on

ASI

C2b

PNS,

bra

in, s

pina

l cor

d,

retin

a, t

aste

cel

lsN

AN

AM

odu

late

the

pro

per

ties

of

othe

r A

SIC

s (m

ainl

y A

SIC

2a a

nd A

SIC

3) [

19,2

9]Pa

in?

ASI

C3

PNS

(incl

udin

g sp

ecia

lized

cut

aneo

us

mec

hano

sens

ory

stru

ctur

es),

ret

ina,

ta

ste

cells

, car

otid

b

od

y, in

ner

ear

, tes

tis,

bon

e, lu

ng e

pith

elia

l ce

lls, v

ascu

lar

smo

oth

mus

cle

cells

6.2

–6

.7A

milo

rid

e [2

2], N

SAID

s (a

spiri

n/d

iclo

fena

c) [6

8],

A-3

1756

7 [6

7],

nafa

mo

stat

mes

ilate

[70]

, di

aryl

amid

ines

[72]

, G

d3+ [7

4], C

d2+ [7

6], P

b2+

[77]

, APE

Tx2

[85]

RFa

mid

e p

epti

des

[52,

90],

ar

achi

don

ic a

cid

[54,

142]

, lac

tate

[1

58],

NO

[143

],

hyp

erto

nici

ty [5

4]

Car

diac

pai

n [4

3],

cuta

neo

us a

cidi

c an

d pr

imar

y in

flam

mat

ory

pain

[54]

, sec

onda

ry m

echa

nica

l hy

per

alg

esia

in m

uscl

e an

d jo

ints

[153

,155

], c

utan

eous

an

d vi

scer

al m

echa

nose

nsat

ion

and

mec

hano

-no

cice

ptio

n [1

26,1

63,1

65],

aci

d se

nsin

g in

ga

stro

eso

phag

eal a

ffer

ents

[161

], m

aint

enan

ce o

f re

tinal

inte

grit

y [1

91],

hea

ring

[192

], t

esto

ster

one

hom

eost

asis

[156

], p

oss

ible

impl

icat

ion

in

chem

otra

nsdu

ctio

n of

low

pH

by

caro

tid

bo

dy

[184

] an

d lo

cal v

ascu

lar

cont

rol i

n m

uscl

e [1

86]

Inhi

biti

on: i

nflam

mat

ion

-re

late

d pa

in, c

hron

ic

mus

cula

r pa

in, a

ngin

a,

arth

ritis

, gas

trit

is,

infla

mm

ator

y an

d no

ninfl

amm

ator

y b

owel

dis

ord

ers

ASI

C4

PNS

(low

exp

ress

ion

),

brai

n, s

pina

l cor

d,

pitu

itar

y gl

and,

ret

ina,

in

ner

ear

, im

mun

e ce

lls

NA

??

May

mo

dula

te t

he p

rop

erti

es o

f ot

her

ASI

Cs

[49]

?

ASI

C: A

cid

-sen

sin

g io

n ch

ann

els;

NA

: Not

ap

plic

able

; PN

S: P

erip

her

al n

ervo

us

syst

em.

Page 4: Current perspectives on acid-sensing ion channels: new advances and therapeutic implications

Expert Rev. Clin. Pharmacol. 3(3), (2010)334

Review Noël, Salinas, Baron, Diochot, Deval & Lingueglia

generates a fast transient current, which, by itself, is large enough to depolarize the membrane and trigger action potentials [52]. In the second mode, sustained and modest acidifications of the extracellular medium (between pH 7.3 and 6.7 in rat isoform), which all remain near the physiological pH, generate a window current, due to the overlap between inactivation and activation of the channel, which produces persistent depolarization [53,54]. In the third mode, more severe sustained acidifications (pH <6.0) activate a plateau current through a specific mechanism, also producing a persistent depolarization [51].

Proton sensing & gatingBased on the 3D structure [40,41], hypotheses have been developed to describe proton sensing and gating of ASIC channels. Several regions have been involved in pH sensing. A cluster of acidic amino acids in the extracellular domain forms an ‘acidic pocket’ with carboxyl–carboxylate pairs [40]. This putative pH sensor may associate binding of protons and opening of the channel via the TM1 domain [40,55]. Another ‘proton-sensitivity signature’ is present in a second region located at the base of the palm domain [56]. Protonation of amino acids that form a ring of negative charges around the outer entrance of the channel in a third region in the wrist domain are probably also necessary for gating [57]. The global conformational change induced by protons binding to the acidic pocket, palm and wrist domains, could open the channel pore through a rotation of the transmembrane helices [40,41,58,59], releasing a constriction formed by the crossing of TM2 domains that occludes the ion conduction pathway (Figure 1) [41]. Na+ ions may access the pore either through three oval-shaped fenestra-tions located at the base of the extracellular domain near the junction with transmembrane domains [40], and by a path along the threefold axis of symmetry in the extracellular domain [41]. Displacement during acidification of the Ca2+ ions bound to the channel in the closed state also seems to be an important feature of ASIC gating [40,57,60–62]. Interestingly, ASICs are regulated by extracellular Ca2+ with a dual inhibitory and stimulatory effect [6,43,57,63–65]. The cytosolic domains, for which the 3D structure is not known, have also been involved in the regulation of channel gating [42,51,66] but the mechanisms remain poorly characterized.

Pharmacology Small molecules & metal ionsAmiloride, a K+-sparing diuretic, is a weak, nonselective inhibitor of ASIC channels in the micromolar concentration range that acts as a pore blocker. Amiloride has an additional paradoxical -enhancing effect on the sustained current component of ASIC3 [22,53]. Amiloride derivatives, benzamil and EIPA, are also reversible, poorly selective blockers of ASIC channels. The ami-dine A-317567 has been described as a more specific inhibitor of ASICs, with IC

50 between 2 and 30 µM on native ASIC currents

in DRG neurons [67]. A-317567 was shown to be more potent than amiloride in vitro and in vivo in models of inflammatory and postoperative pain in rats [67]. ASIC1a and ASIC3 are also directly inhibited by therapeutic concentrations of NSAIDs (IC

50

~92–350 µM) [68,69], and the clinically used protease inhibitor,

nafamostat mesilate, reversibly blocks recombinant human ASIC1a, ASIC2a and ASIC3 channels expressed in Xenopus oocytes (IC

50 ~2.5–71 µM) [70]. The aminoglycosides streptomy-

cin and neomycin partially and reversibly inhibit ASIC currents in DRG neurons (IC

50 ~32–44 µM) [71]. Diarylamidines (e.g.,

DAPI, diminazene, hydroxystilbamidine and pentamidine) have recently been shown to inhibit ASIC currents in primary culture of hippocampal neurons (IC

50 ~2.8, 0.3, 1.5 and 38 µM, respec-

tively), and diminazene reversibly blocks recombinant ASIC1a, ASIC1b, ASIC2a and ASIC3 channels expressed in Chinese ham-ster ovary (CHO) cells [72]. ASIC channels are inhibited by a vari-ety of heavy-metal ions (Gd3+, Pb2+, Ni2+, Cd2+ and Cu2+) [73–77] and bivalent cations (Ca2+ and Zn2+) [6,63,78]. Extracellular Zn2+ at micromolar concentrations also potentiates the acid activation of homomeric and heteromeric ASIC2a-containing channels [79], and the native ASIC currents in rat hippocampal neurons [80].

Peptide toxinsTwo toxins that selectively and efficiently block ASICs in vitro and in vivo in a subtype-specific manner have been isolated from animal venoms [81]. The spider toxin, Psalmotoxin 1 (PcTx1), blocks ASIC1a homomeric channel with an IC

50 of approximately

0.9 nM [82–84]. The sea anemone toxin, APETx2, blocks homo-meric ASIC3 with an IC

50 of approximately 63 nM, as well as

ASIC3-containing heteromeric channels (IC50

~0.1–2 µM) [85,86]. PcTx1 is a gating modifier that binds with the highest affinity to the desensitized state of the channel at a site in the extracel-lular domain close to the acidic pocket [83,87,88]. The toxin causes both a robust shift of the steady-state desensitization curve and a shift of the proton-activation curve to higher pH. The resulting increase of the affinity for protons inhibits ASIC1a by shifting most channels into the desensitized state at pH 7.4 [87]. PcTx1 also interacts with ASIC1b, and the toxin binds to the open state and does not inhibit the channel but promotes its opening at slight acidification [89].

NeuropeptidesFMRFamide and structurally related peptides, such as FRRFamide and neuropeptide FF (NPFF, FLFQPQRFamide) potentiate H+-gated currents of heterologously expressed ASIC1 and ASIC3, but not ASIC2a (EC

50 ~10–50 µM; threshold

~1 µM) [52,90–92]. The neuropeptides increase the peak ampli-tude and/or slow inactivation of the H+-gated current, which induces or increases the sustained phase during acidification. The peptide effect seems to be by direct binding on the chan-nel [52,90,93]. The peptides alter steady-state desensitization of ASIC1a [94]. ASIC2a is capable of increasing the response to peptides in heteromeric channels containing ASIC1a or ASIC3 subunits [91,95]. The modulation of ASICs by RFamide-related neuropeptides is strongly pH dependent (half-maximal effect achieved at pH 5.6–6.0 [96]) and requires FMRFamide addition at pH 7.4 (i.e., when the channel is closed) [90]. Knockout mice have demonstrated the major role of ASIC3 in mediating the sensory response to FMRFamide and FRRFamide [93,97]. The ASIC1 contribution to the effect seems to be very modest [93,97],

Page 5: Current perspectives on acid-sensing ion channels: new advances and therapeutic implications

www.expert-reviews.com 335

ReviewAcid-sensing ion channels: new advances & therapeutic implications

which is consistent with the lower potency of FMRFamide and related peptides in modulating ASIC1a compared with ASIC3. Two endogeneous opioid neuropeptides, dynorphin A and big dynorphin, have recently been shown to directly interact with ASIC1a and enhance channel activity by preventing its steady-state desensitization [98].

ASICs in the brain Native ASIC-like currents have been reported in almost all types of brain neurons. In rodents, ASIC1a and ASIC2, but not ASIC3, are widely distributed in the cortex, striatum, hip-pocampus, olfactory bulb, amygdala and cerebellum [99,100]. ASIC1 and ASIC2 are detected in brain synaptosomal fractions, suggesting a synaptic localization [33,101,102]. Genetic deletion of ASIC1 eliminates the H+-evoked currents in hippocampal [95,101] and amygdala neurons [99], which illustrates the central importance of this isoform in the brain. ASIC channels in the brain have been largely studied with ASIC1-knockout mice, revealing physiological and pathophysiological roles in different brain regions (Figure 2).

ASICs in the hippocampus: dendrite morphology, synaptic plasticity, learning & memoryFunctional ASIC channels have been detected in cultured hippo campal neurons [80,101,103]. The important function of ASIC1 in the hippocampus is demonstrated in ASIC1-knockout mice, which show impaired hippocampal long-term potentia-tion and defects in spatial learning and memory [101]. In hip-pocampal neurons, ASIC1a and ASIC2a are present in the postsynaptic membrane of the dendritic tree and dendritic spines [27,33,101]. The presence of ASIC2a increases the density of ASIC1a in spines through an association with the PDZ-containing protein PSD-95 [33]. It has been proposed that the secondary increase in intracellular Ca2+ mediated by ASIC1a positively affects the spine density in hippocampal neurons through the Ca2+-mediated CaMKII signaling pathway [48]. However, ASIC1-knockout mice do not have reduced spine density [48]. The contribution of ASIC1a to the presynaptic function is not clear since paired pulse facilitation (an index of presynaptic activity and neurotransmitter-release probability) is normal in hippocampal slices from ASIC1-knockout mice [101], but a recent study has shown an increase of the probability of neuro- transmitter release in cultured hippocampal neurons from the same knockout mice [104].

ASIC1a in amygdala: fear, chemosensing & depressionExpression of ASIC1a is particularly high in the amygdala, a brain region involved in fear, arousal and emotions [99]. In this structure, ASIC1a seems to contribute to fear responses to a variety of aversive stimuli, perhaps, in part, through a contribution to synaptic plastic-ity underlying the acquisition of conditioned fear [101,105]. ASIC1-knockout mice show impaired fear conditioning [99] and innate fear [106], and restoring ASIC1a expression in the basolateral amy-gdala of knockout mice rescues contextual fear conditioning [107]. Conversely, overexpression of ASIC1a in transgenic mice increases

fear conditioning [108]. Recently, ASIC1a has been shown to act in the amygdala as a pH sensor that contributes to the production of fear behavior associated with the inhalation of CO

2 and acidosis [109].

Loss of ASIC1 in knockout mice or pharmacological inhi-bition of homomeric ASIC1a channels by the PcTx1 toxin, or of ASIC channels by A-317567, generates antidepressant effects [110]. Restoring ASIC1a in the amygdala with a viral vector eliminates the antidepressant-like phenotype in ASIC1a-knockout mice, suggesting that the amygdala plays a key role in the anti-depressant-like effects associated with ASIC1a deletion or inhibition [110].

ASIC1a & neurodegeneration Activation of ASIC1a during the metabolic acidosis accompanying experimental stroke has been proposed to contribute to neuronal death associated with brain ischemia [38,46]. ASIC1a has also been proposed to contribute to axonal degeneration in autoimmune inflammation of the CNS in a mouse model of multiple scle-rosis, where significant tissue acidosis occurs [111]. Interestingly, amiloride and, to a lesser extent, spider venom containing the PcTx1 toxin, are neuroprotective in a 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine model of Parkinson’s disease [112]. Amiloride and benzamil also improve pathological changes in cellular and animal models of Huntington’s disease [113], and amiloride reduces hippocampal neuronal degeneration following traumatic brain injury in rats [114].

ASIC1a & seizure It has long been known that breathing carbon dioxide makes brain tissue more acidic and stops seizures, and seizures them-selves are known to reduce brain pH. The anti-epileptic effect of low pH has been ascribed to an increase in the inhibitory tone of inhibitory interneurons through ASIC1a activation [115]. Interestingly, ASIC currents have been reported to be stronger in inter neurons than in excitatory neurons [116]. ASIC1a activa-tion can, therefore, have an inhibitory effect and a protective function in the brain. Whether ASIC channels are protective or deleterious may depend on the location of ASIC activation, the magnitude and duration of acidosis, and the presence of modulators of ASIC function [115]. Several recent studies also have shown an amiloride inhibition of seizures in rats and mice [117–119], raising the additional possibility of some contribution of ASICs to the generation of seizure.

Roles at the synapseThe localization of ASIC1a and the phenotype of knockout mice suggest a role at the synapse. A proton pump acidifies synaptic vesi-cles [120] and provides the electrochemical gradient for transmitter uptake. When vesicles fuse with the plasma membrane, protons are released along with neurotransmitters into the synaptic cleft, which may modulate synaptic transmission. In mammalian cone photoreceptor ribbon synapses, discharged protons can affect the presynaptic Ca2+ channel and vesicle release [121,122]. Similarly, H+ released in the cleft might activate synaptic ASICs in conven-tional synapses. This post synaptic ASIC-generated depolarization

Page 6: Current perspectives on acid-sensing ion channels: new advances and therapeutic implications

Expert Rev. Clin. Pharmacol. 3(3), (2010)336

Review

TongueSour taste (ASIC2a)

JointsSecondary mechanical hyperalgesia (ASIC3)

BrainSynaptic transmission and plasticityLearning and memoryInnate and conditioned fearChemosensing of the amygdala (ASIC1a)

Gastrointestinal tractAcid sensing (ASIC3)Mechanosensation (ASIC1, 2 and 3)Mechano-nociception (ASIC3)

Auditory systemHearing (ASIC3)Noise susceptibility (ASIC2a)

Carotid bodyChemosensation (ASIC1 and/or 3)

Heart and aortic archCardiac pain (ASIC3)Baroreceptor reflex (ASIC2a)

RetinaVisual transduction (ASIC1a, 2a and 3)Maintenance of retinal integrity (ASIC3)

MusclePrimary hyperalgesia (ASIC1a)Secondary mechanical hyperalgesia (ASIC3)Local vascular control (ASIC3)

SkinAcidic and primary inflammatory pain (ASIC3)Mechanosensation (ASIC2a and 3)

Spinal cordCentral sensitization and modulation of pain (ASIC1a)

Noël, Salinas, Baron, Diochot, Deval & Lingueglia

has been proposed to facilitate the release of the Mg2+ block of the NMDA receptor, and contributes to synaptic plasticity [101]. However, ASIC-mediated postsynaptic currents have not been detected during synaptic transmission [101,102,104,123], and the extent of the synaptic pH changes and whether or not these changes are sufficient to activate ASIC1a, remains to be determined. In addition, and as mentioned earlier, it is conceivable that ASIC1a could also contribute to the presynaptic function. Interestingly, it was recently demonstrated that protons can serve as intercellular transmitters

between intestinal cells and defecation muscles in C. elegans [124].

ASICs & nociceptionAcid-sensing ion channels are present in primary sensory neurons of the trigeminal, vagal and dorsal root ganglia. ASIC1, ASIC2 and ASIC3 are expressed in small and medium neurons (i.e., the nociceptive neurons that are able to detect noxious stimuli) [68,125–130]. ASIC2 and ASIC3 have also been found in large non-nociceptive neurons that mostly correspond to low-threshold mechanoreceptors. The proteins have been detected in the soma and in the peripheral nerve endings of DRG neurons [125–128]. The presence of ASICs on central projections in the dorsal horn of the spinal cord is less clear [125,131,132]. ASICs support most of the native proton-activated cation currents in sensory neurons, although a significant part of the sustained response to low pH can be attributed to the capsaicin receptor TRPV1 [133]. The native ASIC-like responses are approximately tenfold more sensitive to changes in H+ than the TRPV1 responses.

In rats, activation of nociceptive C-fibers by acid is inhibited by amiloride [134], and ASIC-mediated depolarization of sensory neurons triggers action potentials upon extracellular acidifi-cation [52,135]. ASIC3 is a remarkable subunit because, in addition to its large expression in sensory neurons and steep sensitivity to protons, it can adjust its response to a wide range of extra cellular pH variations by generating transient and sustained currents with a significant impact on the neuronal excitability. As previously described, ASIC3 has three distinct modes of pH sensing. The two persistent modes of activity might explain how ASIC3 underlies a sustained sensation of pain triggered by modest acidifications near the physiological pH (between 7.4 and 6.5) or by more severe decreases in pH, even if its fast transient current component inac-tivates rapidly [53,54]. Therefore, ASICs have been proposed to be sensors of acidic pH on the peripheral terminals of primary sensory neurons and to participate in the perception of pain that accompa-nies tissue acidosis in conditions such as ischemia, inflammation, tumors, infections, lesions or traumatic injuries. Several factors may participate in the drop of extra cellular pH associated with these conditions, such as release of the acidic content of lyzed cells, degranulation of mast cells, organic acids released by metabolism or infectious agents, such as lactic acid released by ischemic muscle and heart, as well as damage to the mucosal barrier of the GI tract or damage to the urinary tract epithelium, exposing the underlying tissue to gastric acid and acidic urine, respectively.

ASICs in somatic painASICs in cutaneous painThe role of ASICs in cutaneous nociception has remained con-troversial until recently because of the apparent lack of phenotype and the conflicting results generated from studies of knockout mice for ASIC1, ASIC2 and ASIC3 [126,136,137] and from a study of ASIC3 dominant-negative mice [138]. For instance, normal, decreased or increased pain behaviors were reported for ASIC3-knockout or -transgenic mice in various studies and different pain models [126,136–138]. However, peripheral injections of the APETx2 toxin (a specific blocker of ASIC3) and knockdown of ASIC3 by

Figure 2. Proposed physiological functions of the different acid-sensing ion channels. ASIC: Acid-sensing ion channels

Page 7: Current perspectives on acid-sensing ion channels: new advances and therapeutic implications

www.expert-reviews.com 337

ReviewAcid-sensing ion channels: new advances & therapeutic implications

intrathecal injections of siRNAs have clearly demonstrated in rat the essential role played by ASIC3 as a sensor of acidic pain, as well as an integrator of molecular signals produced during inflam-mation, where it contributes to primary thermal hyper algesia [54]. Consistent with these data, local peripheral inhibition of ASICs by amiloride and benzamil reduces cutaneous inflammatory pain in rats [139]. This is in agreement with the increased expression of ASICs in rat DRG neurons induced by chronic inflammation of the hind paw [68,135,140,141]. The increase in expression, together with the upregulation of ASIC3 activity by several components of the ‘inflammatory soup’, such as bradykinin, 5-hydroxytryp-tamine [29], hypertonicity [54], arachidonic acid [54,73,142] and nitric oxide [143], may be important for the sensitization of cutaneous nociceptors during inflammation. Interestingly, and consistent with the data in rats, a decrease in pH in the skin of human volunteers has been associated with nonadapting pain [144], and this cutaneous acid-induced pain appears to be largely mediated by ASIC channels, especially at moderate pH (>6.0) [145–147]. A recent study in ASIC3-knockout mice has looked at the nocic-eptive behaviors in both the acute (a few hours) and subacute (1–2 days) phases of inflammation, and reveals an involvement of ASIC3 in the maintenance of primary hyperalgesia (i.e., the increased response to noxious stimuli at the site of injury) in the subacute but not the acute phase [148]. Behavioral responses to cutaneous mechanical and thermal stimuli are unaffected in ASIC1-knockout mice [137], and subcutaneous injections of the PcTx1 toxin in wild-type animals to specifically block homomeric ASIC1a channels have no effect in acute pain models [149], or in a model of inflammatory pain in rats [132]. Consequently, homo-meric ASIC1a channels do not seem to significantly contribute to acute or inflammatory pain at the periphery. However, one cannot exclude a role for ASIC1a in peripheral nociceptor terminals when associated with other ASIC subunits in heteromeric channels.

The contrasting results found in mice and rats could seem sur-prising but may be explained by compensatory mechanisms in knockout animals and/or by species differences in expression of ASIC channels. Indeed, approximately 65% of rat nociceptors innervating the skin express ASIC currents [54], and transient proton-induced ASIC-like currents are less frequent and have smaller current density in the mouse [150].

ASICs in muscle & deep somatic pain Acid-sensing ion channels are expressed in sensory neurons that innervate muscle, joints and bone. ASIC3 is expressed in more than 50% of small muscle afferents in rats [129] and in more than 30% of DRG neurons innervating the knee joint in mice [151]. ASIC expression in DRG is increased in mouse models of mus-cle inflammation [152] and acute arthritis [151]. ASIC3 plays an important role in the generation of secondary mechanical hyper-algesia (i.e., the increased response to noxious stimuli outside the site of injury) and central sensitization achieved in a mouse model of noninflammatory muscular pain induced by repeated acid injections into the muscle [126,153]. ASIC3 is also involved in the development of cutaneous mechanical, but not heat, hyper-algesia induced by muscle inflammation [154]. Conversely, ASIC1

does not participate in acid- or inflammation-induced secondary mechanical hyperalgesia in mouse muscle [152,153], but contributes to primary hyperalgesia in muscle inflammation [152].

Acid-sensing ion channels3 has also been involved in the sec-ondary, but not the primary, mechanical hyperalgesia produced by joint inflammation in mice [155]. Enhanced muscle fatigue also occurs in ASIC3-knockout mice, but only in males, which may be related to lower plasma levels of testosterone [156].

ASICs in visceral pain & mechanosensationAcid-sensing ion channels 3 in cardiac afferents has been pro-posed to be the major detector of myocardial acidity that trig-gers angina during cardiac ischemia [43,157]. Interestingly, organic compounds released during ischemic situations in the heart, but also in muscle and the brain, such as lactate and arachidonic acid, are able to increase ASIC currents in sensory and central neurons [54,73,142,158].

Acid-sensing ion channels transduce acid sensation in gastric sensory nerve endings. ASIC-like currents have been recorded in sensory neurons that innervate the stomach from both dorsal root ganglia and nodose ganglia, and these currents are sensitized by gastric ulceration [159]. ASIC3 has been proposed to contribute to acid sensation in gastroesophageal afferents [160]. ASIC3 does not seem to participate to the acid sensing of vagal afferent neurons in the normal stomach but plays a major role in the increased response to acid associated with experimental gastritis [161]. ASIC3 could, therefore, be important in acid-related, inflammation and ischemia-induced disturbances of gut function and sensation [162], as well as in dyspepsia and GI reflux disease affecting the proximal gastrointesti-nal tract. On the other hand, ASIC2-knockout mice have increased acid-evoked afferent input from the stomach to the brain stem [161].

In addition, ASICs play a role in gastrointestinal mechano-sensory function. Mice lacking ASIC1, 2 and 3 demonstrate altered colonic mechanosensory function. Disruption of ASIC1a increases the mechanical sensitivity in all colonic and gastro-esophageal sensory afferents [137]. Disruption of ASIC2 has mixed effects, while in ASIC3-knockout mice, all afferent classes except gastro esophageal mucosal afferents have markedly reduced mech-anosensitivity [160,163]. Quantitive PCR of laser-captured neurons has identified ASIC3 as the most abundant ASIC transcript in colonic DRG neurons [164], consistent with its contribution to the perception of noxious mechanical stimuli in the mouse colon [165]. ASIC3 is also required for the sensitization (but not the activation) by the acidic inflammatory soup of colon afferent fibers to stretch [165]. It is important to mention that mechanical hypersensitiv-ity of the colon underlies, in part, the chronic abdominal pain experienced by many patients with bowel disorders. Increased ASIC3 expression has been detected in the colonic mucosa of patients with Crohn’s disease (an inflammatory bowel disease) [140], and ASIC3 contributes to the visceral nociceptive behavior in a mouse model of noninflammatory bowel disorder, mimicking the pathophysio logy of irritable bowel syndrome [166].

Acid-sensing ion channel-like currents have been recorded on rat vagal pulmonary sensory neurons [167], and ASIC3 is expressed in rat vagal and glossopharyngeal sensory ganglia [168], as well as in

Page 8: Current perspectives on acid-sensing ion channels: new advances and therapeutic implications

Expert Rev. Clin. Pharmacol. 3(3), (2010)338

Review Noël, Salinas, Baron, Diochot, Deval & Lingueglia

spinal afferent neurons projecting to the rat lung and pleura [169]. ASICs have been proposed to play a role in the effect of acidosis on airway basal tone and responsiveness in the guinea pig. [170]. ASICs could, therefore, participate in the response to airway acidification, such as cough and bronchoconstriction [171–173], which are associ-ated with both physiological (e.g., exercise) and pathophysiological (e.g., chronic obstructive pulmonary disease) conditions.

Central ASIC1a & pain modulationAcid-sensing ion channels 1a and ASIC2 are largely expressed in spinal second-order sensory neurons that receive primary afferent inputs. Their expression is upregulated during peripheral inflam-mation, and they may play a role in the processing of noxious stimuli and in the generation of hyperalgesia and allodynia in persistent pain [50,132,174]. Recently, ASIC1a has been involved in central sensitization associated with prolonged pain [132]. FMRFamide-related peptides, which directly modulate ASIC1- and ASIC3-containing channels, are also increased in the spinal cord during chronic inflammation [175–177], which could modulate the response to noxious acidosis. ASIC1-knockout mice do not exhibit sensory deficits [99,101]. However, a role for central ASIC1a in the pain pathway has been clearly demonstrated by intrathecal and intracerebroventricular injections of the PcTx1 toxin in mice and rats, which have a potent analgesic effect in several models of acute and chronic pain [149]. Knockdown of the channel after intrathecal injection of antisense oligonucleotides has a similar analgesic effect. Blocking of the ASIC1a homomeric channels at the spinal and/or supraspinal level results in an activation of the endogenous enkephalin pathway and in increased levels of Met-enkephalin in the cerebrospinal fluid [149], producing strong analgesic effects.

ASICs & other sensory modalities ASICs & mechanotransductionAcid-sensing ion channels are expressed in small DRG neurons, including high-threshold mechanoreceptors, and ASIC2a and ASIC3 are also expressed, as mentioned earlier, in large DRG neu-rons that comprise low-threshold mechanoreceptors. The ASIC2 and ASIC3 proteins have been detected in specialized cutaneous mechanosensory structures, such as Meissner corpuscles, Merkel nerve endings and palisades of lanceolate nerve endings surround-ing the hair shaft [125,126,128]. Interestingly, genetic dissection of mechanosensory transduction in C. elegans have identified mem-bers of the ENaC/DEG/ASIC family of ion channels (i.e., the degenerin MEC-4 and MEC-10), as mechanotransducers in body touch neurons [10–13,178–180]. Thus, ASICs appear to be good can-didates for mechanosensitive ion channels in mammals. However, genetic studies in mice only partially support a role in cutaneous mechanotransduction. Some studies of knockout mice for ASIC2 or ASIC3, but not ASIC1, have reported subtle alterations in cuta-neous mechanical sensitivity (slight increase or slight decrease, depending on the modality tested and the genetic background) [126,128,137,181], while other studies have failed to confirm such a role [103,182]. Very recently, ASIC2 has been detected in the soma and sensory endings of nodose ganglia aortic baroreceptor neurons,

and knockout mice show an impaired baroreceptor reflex, sug-gesting a role for ASIC2, maybe as a pressure sensor, in barorecep-tor sensitivity and neural control of blood pressure [183]. ASIC1, ASIC2 and ASIC3 also differentially contribute to gastrointesti-nal mechanosensory function. Available data so far do not actually support a direct role for ASICs as mechanically gated channels, and their participation in mechanotransduction is more likely to be indirect.

ASICs & chemotransductionAcid-sensing ion channels 1 and ASIC3 are expressed in rat carotid body, and extracellular acidosis evokes ASIC-like inward current in glomus cells – the chemosensitive cells that monitor arterial blood oxygen, carbon dioxide, and pH to regulate breathing – suggesting a contribution of ASIC channels to the detection of low blood pH by the carotid body [184].

A role for ASIC channels in the local vascular control has also been proposed [123]. ASIC3-expressing sensory nerve endings have been shown to innervate muscle arterioles [129]. These afferents may correspond to muscle metaboreceptors that detect changes in muscle metabolism, consistent with the contribution of ASICs to the exercise pressor reflex evoked by metabolic but not mechanical signals generated by contracting skeletal muscle [185,186].

ASICs in vision, hearing & tasteMost ASICs have been detected in rodent retina [187–191]. ASIC2 is a negative modulator of rod phototransduction [188], while ASIC1a is a positive modulator of cone phototransduction and adapta-tion [190]. Inactivation of the ASIC2 gene in the mouse also sen-sitizes the retina toward light-induced degeneration [188]. ASIC3 is also expressed in the rod inner segment of photoreceptors, in horizontal cells, in some amacrine cells and in retinal ganglion cells. Inactivation of ASIC3 enhances visual transduction in mice 2–3 months of age but induces late-onset rod photorecep-tor degeneration and death in older mice, suggesting an important role of ASIC3 in the maintenance of retinal integrity [191].

Acid-sensing ion channels 1, 2 and 3 are expressed in the peripheral auditory system [192–194], and currents corresponding to homomeric ASIC1a- and ASIC2a-containing channels, but not ASIC3 channels, have been recorded in the inferior colliculus of the central auditory system [195]. ASIC2 and ASIC3 are mainly expressed in spiral ganglion neurons of the adult cochlea, which transmit primary acoustic information from cochlear hair cells to the brain. ASIC2-mutant mice show no hearing loss [103,193] but have altered noise susceptibility [193]. ASIC3-knockout mice have normal hearing at 2 months, but develop progressive hearing loss with age [192]. Interestingly, ASIC3-dependent hearing loss in mothers leads to inadequate maternal response to pups’ calls, affecting the social development of pups in adolescence [196].

Several ASICs have been localized to taste buds, with some species-specific expression [197–201]. ASIC2-knockout mice have no significant impairment of responses to sour stimuli [199], but a recent study with two patients unresponsive to sour stimuli but not to other taste modalities, suggests an implication of ASIC channels in sour-taste perception in the human tongue [202].

Page 9: Current perspectives on acid-sensing ion channels: new advances and therapeutic implications

www.expert-reviews.com 339

ReviewAcid-sensing ion channels: new advances & therapeutic implications

Other roles for ASICsAcid-sensing ion channels have been detected in the testis (hASIC3; note that ASIC3-knockout males display lower plasma levels of testosterone [156]) [20,23], pituitary gland (ASIC4) [25], lung epithelial cells (ASIC1a and 3) [203,204], immune cells (ASIC1, 3 and 4) [111], urothelial cells [205], adipose cells (ASIC3) [206] and bone (ASIC1–3) [207,208]. In non-neuronal joint cells, ASIC3 regu-lates pH responsiveness and controls release of the extracellular matrix polysaccharide hyaluronan [209].

Acid-sensing ion channels are expressed in vascular smooth muscle cells (ASIC1–3) [210–213], and ASIC2 has been involved in vascular smooth muscle cell migration [211]. ASIC2-knockout mice also have defects in pressure-induced vasoconstriction in mouse middle-cerebral arteries [213]. A role for ASIC2 in migration of glioma cells has also been proposed [214].

Genetic associations with pathologiesNo inherited human disease caused by mutations in ASICs has been identified so far. However, recent studies suggest genetic associations between the ASIC2 locus and multiple sclerosis, autism and mental retardation [215–217], and a possible link, although marginally significant, between the ASIC1 gene and major depression [218].

Expert commentaryAcid-sensing ion channels constitute a relatively recently iden-tified family of excitatory depolarizing cation channels. Their only known activators are extracellular protons, with a very strong dependence on pH. The structure and function of these channels remain partially understood, but an important step forward came from the recent determination of the 3D struc-ture of chicken ASIC1, showing trimeric organization. ASICs are broadly expressed in the nervous system, both in central and sensory neurons. Available data suggest an essential role for the ASIC1a and ASIC3 isoforms in the central and the periph-eral nervous system, respectively. Nonsubtype-specific inhibi-tors of ASICs are available, such as the nonselective, clinically used diuretic amiloride and A-317567, another small molecule experimentally developed by Abbott Laboratories (UK), which appears to be more selective than amiloride but also exhibits relatively low potency. In addition to these nonselective and/or nondiscriminative inhibitors, more potent and subtype-selective blockers are available, such as a spider peptide toxin (PcTx1) and a sea anemone peptide toxin (APETx2), which specifically block the ASIC1a- and ASIC3-containing channels, respectively. Use of these pharmacological inhibitors has shown, in association with genetic studies using knockout mice of various isoforms, a clear implication of ASICs in brain physiology and patho-physiology, as well as in pain perception and modulation. Of particular interest is the role of ASIC1a in the amygdala, where it contributes to innate and acquired fear, chemosensing of carbon dioxide and depression, making inhibitors of this channel new potential pharmaceutical tools in the management of psychiat-ric disorders, including anxiety, panic and depression. ASIC1a could also represent an interesting target to reduce neuronal

degeneration associated with brain acidosis in stroke and neu-rodegenerative diseases. On the other hand, ASIC1a activators could be beneficial in modulating seizure termination. Central inhibitors of ASIC1a could also offer a new potential line of treatment for chronic pain through positive modulation of the endogenous opioid system. ASIC3 also emerges as a promis-ing target for treatment of persistent somatic and visceral pain, including inflammation-related pain, chronic muscular pain, angina, arthritis, gastritis and inflammatory and noninflam-matory bowel disorders. Finally, implication of ASIC channels in retinal function, carotid body chemotransduction, baro-receptor reflex, hearing, sour-taste perception, and probably in other functions yet to be identified, may also lead to interesting therapeutical applications and must be taken into consideration in any ASIC therapy protocol to minimize adverse side effects.

Five-year viewProtons are the only known activators of ASICs and it is rea-sonable to assume that ASIC functions are supported by their capacity to behave as extracellular pH sensors. However, the possibility of the existence of other stimuli, such as neuro-peptides, cannot currently be ruled out, and some of the func-tions assigned to ASICs may be unrelated to the detection of extracellular pH.

Acid-sensing ion channels are almost ubiquitous in mam-malian CNS neurons. They have been implicated in a number of physiological and pathophysiological conditions but basic knowledge of their contribution to neuronal function remains limited. For instance, the interesting possibility of a signaling role for protons at the synapse via ASICs definitely requires further investigation. Little is also known about the amplitude of pH fluctuations in the brain during normal physiological processes, as well as in pathological situations. Similarly, it is necessary to extend our knowledge of the implication of ASICs in pain, as well as the precise mechanisms of action underlying their roles in nociception, to develop effective therapeutical strategies involving these channels. In this respect, progress in our knowledge of the 3D structure and further exploration of the molecular dynamics of ASICs will greatly facilitate the dis-covery and optimization of new clinically relevant drugs target-ing these channels. Little is still known about ASIC regulation, and there is a need for further investigation in this area, con-sidering the participation of these channels in highly regulated processes in the central and peripheral nervous systems. The molecular basis of the implication of ASICs in mechanosensory function remains largely unknown. It is important to determine if, in certain cellular contexts for instance, some ASICs could be able to transduce mechanical signals, perhaps with the help of accessory proteins present on both side of the plasma membrane, as shown for the related mechanosensitive degenerin channels in C. elegans.

Acid-sensing ion channels are broadly expressed in the nerv-ous system, but there is emerging evidence of expression in non -neuronal cells and it is probable that novel roles outside the nervous system, in relation or not with their ability to

Page 10: Current perspectives on acid-sensing ion channels: new advances and therapeutic implications

Expert Rev. Clin. Pharmacol. 3(3), (2010)340

Review Noël, Salinas, Baron, Diochot, Deval & Lingueglia

sense extracellular pH, will be assigned to these channels. For instance, the function of mammalian ASIC4, which is not acti-vated by protons, remains poorly characterized. The multigenic nature of the ASIC family also makes the analysis of multiple ASIC-knockout mice probably necessary to clarify and expand their repertoire of functions in physiological and pathophysi-ological situations both inside and outside the nervous system. Species-specific differences in expression also need to be taken in consideration when evaluating these functions.

Financial & competing interests disclosureWe thank the Association Française contre les Myopathies (AFM), the Agence Nationale de la Recherche (ANR), the Association pour la Recherche sur le Cancer (ARC), the Institut National du Cancer (INCa), and the Institut UPSA de la Douleur (IUD) for financial support. The authors have no other relevant affiliations or financial involvement with any organization or entity with a financial interest in or financial conflict with the subject matter or materials discussed in the manuscript apart from those disclosed.

No writing assistance was utilized in the production of this manuscript.

References1 Price MP, Snyder PM, Welsh MJ. Cloning

and expression of a novel human brain Na+ channel. J. Biol. Chem. 271(14), 7879–7882 (1996).

2 Waldmann R, Champigny G, Voilley N, Lauritzen I, Lazdunski M. The mammalian degenerin MDEG, an amiloride-sensitive cation channel activated by mutations causing neurodegeneration in Caenorhabditis elegans. J. Biol. Chem. 271(18), 10433–10436 (1996).

3 Garcia-Anoveros J, Derfler B, Neville-Golden J, Hyman BT, Corey DP. BNaC1 and BNaC2 constitute a new family of human neuronal sodium channels related to degenerins and epithelial sodium channels. Proc. Natl Acad. Sci. USA 94(4), 1459–1464 (1997).

4 Krishtal OA, Pidoplichko VI. A receptor for protons in the nerve cell membrane. Neuroscience 5(12), 2325–2327 (1980).

5 Gruol DL, Barker JL, Huang LY, MacDonald JF, Smith TG Jr. Hydrogen ions have multiple effects on the excitability of cultured mammalian neurons. Brain Res. 183(1), 247–252 (1980).

6 Waldmann R, Champigny G, Bassilana F, Heurteaux C, Lazdunski M. A proton-gated cation channel involved in acid-sensing. Nature 386(6621), 173–177 (1997).

7 Kellenberger S, Schild L. Epithelial sodium channel/degenerin family of ion channels: a variety of functions for a shared structure. Physiol. Rev. 82(3), 735–767 (2002).

8 Canessa CM, Horisberger JD, Rossier BC. Epithelial sodium channel related to proteins involved in neurodegeneration. Nature 361(6411), 467–470 (1993).

9 Lingueglia E, Voilley N, Waldmann R, Lazdunski M, Barbry P. Expression cloning of an epithelial amiloride-sensitive Na+

channel. A new channel type with homologies to Caenorhabditis elegans degenerins. FEBS Lett. 318(1), 95–99 (1993).

10 Driscoll M, Chalfie M. The mec-4 gene is a member of a family of Caenorhabditis elegans genes that can mutate to induce neuronal degeneration. Nature 349(6310), 588–593 (1991).

11 Huang M, Chalfie M. Gene interactions affecting mechanosensory transduction in Caenorhabditis elegans. Nature 367(6462), 467–470 (1994).

12 Liu J, Schrank B, Waterston RH. Interaction between a putative mechanosensory membrane channel and a collagen. Science 273(5273), 361–364 (1996).

13 Syntichaki P, Tavernarakis N. Genetic models of mechanotransduction: the nematode Caenorhabditis elegans. Physiol. Rev. 84(4), 1097–1153 (2004).

Key issues

• Acid-sensing ion channels (ASICs) form a family of voltage-independent, excitatory cation channels activated by extracellular protons. They predominantly conduct Na+ and show a very strong dependence on pH.

• ASICs are broadly expressed in the nervous system. Almost all isoforms are present in sensory neurons, while only ASIC1a and ASIC2 have been detected in central neurons in rodents.

• ASICs are nonselectively blocked by amiloride, more selectively blocked by the amidine A-317567, and specifically blocked by two peptide toxins, the spider toxin PcTx1, which inhibits ASIC1a channels, and the sea anemone toxin APETx2, which inhibits ASIC3-containing channels.

• Genetic and pharmacologic data in mice and rats demonstrate a central role for the isoform ASIC1a in the CNS including synaptic plasticity, learning and memory in the hippocampus, and fear, chemosensing and depression in the amygdala.

• ASIC1a has been involved in neuronal degeneration associated with brain acidosis in stroke and neurodegenerative diseases in mice.

• ASIC1a contributes to seizure termination by CO2 and low pH

in mice.

• ASIC1a participates in the central modulation of pain through an interaction with the endogenous opioid system (inhibition of ASIC1a activates the endogenous enkephalin pathway).

• Genetic and pharmacologic data in rodents demonstrate a crucial role for the isoform ASIC3 in the peripheral nervous system in somatic and visceral pain, including inflammation-related pain, chronic muscular pain, angina, arthritis and gastritis, as well as inflammatory and noninflammatory bowel disorders.

• ASICs have also been involved in rodents in mechanosensation, retinal function, baroreceptor reflex, carotid body chemotransduction, hearing and sour-taste perception, and have been detected in many other tissues (e.g., testis, vascular smooth muscle cells, pituitary gland, epithelial cells, immune cells and bone).

• Inhibitors of ASICs may, therefore, constitute new therapeutic tools in the management of psychiatric disorders, neurodegenerative diseases (inhibitors of ASIC1a), and pain (inhibitors of ASIC1a and/or ASIC3), while activators of ASIC1a could be beneficial for seizure treatment.

Page 11: Current perspectives on acid-sensing ion channels: new advances and therapeutic implications

www.expert-reviews.com 341

ReviewAcid-sensing ion channels: new advances & therapeutic implications

14 Lingueglia E, Champigny G, Lazdunski M, Barbry P. Cloning of the amiloride-sensitive FMRFamide peptide-gated sodium channel. Nature 378(6558), 730–733 (1995).

15 Lingueglia E, Deval E, Lazdunski M. FMRFamide-gated sodium channel and ASIC channels: a new class of ionotropic receptors for FMRFamide and related peptides. Peptides 27(5), 1138–1152 (2006).

16 Chen CC, England S, Akopian AN, Wood JN. A sensory neuron-specific, proton-gated ion channel. Proc. Natl Acad. Sci. USA 95(17), 10240–10245 (1998).

17 Bassler EL, Ngo-Anh TJ, Geisler HS, Ruppersberg JP, Grunder S. Molecular and functional characterization of acid-sensing ion channel (ASIC) 1b. J. Biol. Chem. 276(36), 33782–33787 (2001).

18 Champigny G, Voilley N, Waldmann R, Lazdunski M. Mutations causing neurodegeneration in Caenorhabditis elegans drastically alter the pH sensitivity and inactivation of the mammalian H+-gated Na+ channel MDEG1. J. Biol. Chem. 273(25), 15418–15422 (1998).

19 Lingueglia E, de Weille JR, Bassilana F et al. A modulatory subunit of acid sensing ion channels in brain and dorsal root ganglion cells. J. Biol. Chem. 272(47), 29778–29783 (1997).

20 Babinski K, Le KT, Seguela P. Molecular cloning and regional distribution of a human proton receptor subunit with biphasic functional properties. J. Neurochem. 72(1), 51–57 (1999).

21 de Weille JR, Bassilana F, Lazdunski M, Waldmann R. Identification, functional expression and chromosomal localisation of a sustained human proton-gated cation channel. FEBS Lett. 433(3), 257–260 (1998).

22 Waldmann R, Bassilana F, De Weille JR, Champigny G, Heurteaux C, Lazdunski M. Molecular cloning of a non-inactivating proton-gated Na+ channel specific for sensory neurons. J. Biol. Chem. 272, 20975–20978 (1997).

23 Ishibashi K, Marumo F. Molecular cloning of a DEG/ENaC sodium channel cDNA from human testis. Biochem. Biophys. Res. Commun. 245(2), 589–593 (1998).

24 Akopian AN, Chen CC, Ding Y, Cesare P, Wood JN. A new member of the acid-sensing ion channel family. Neuroreport 11(10), 2217–2222 (2000).

25 Grunder S, Geissler HS, Bassler EL, Ruppersberg JP. A new member of acid-sensing ion channels from pituitary gland. Neuroreport 11(8), 1607–1611 (2000).

26 Duggan A, Garcia-Anoveros J, Corey DP. 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(7), 5203–5208 (2002).

27 Hruska-Hageman AM, Wemmie JA, Price MP, Welsh MJ. Interaction of the synaptic protein PICK1 (protein interacting with C kinase 1) with the non-voltage gated sodium channels BNC1 (brain Na+ channel 1) and ASIC (acid-sensing ion channel). Biochem. J. 361(Pt 3), 443–450 (2002).

28 Baron A, Deval E, Salinas M, Lingueglia E, Voilley N, Lazdunski M. Protein kinase C stimulates the acid-sensing ion channel ASIC2a via the PDZ domain-containing protein PICK1. J. Biol. Chem. 277(52), 50463–50468 (2002).

29 Deval E, Salinas M, Baron A, Lingueglia E, Lazdunski M. ASIC2b-dependent regulation of ASIC3, an essential acid-sensing ion channel subunit in sensory neurons via the partner protein PICK-1. J. Biol. Chem. 279(19), 19531–19539 (2004).

30 Anzai N, Deval E, Schaefer L, Friend V, Lazdunski M, Lingueglia E. The multivalent PDZ domain-containing protein CIPP is a partner of acid-sensing ion channel 3 in sensory neurons. J. Biol. Chem. 277(19), 16655–16661 (2002).

31 Deval E, Friend V, Thirant C et al. Regulation of sensory neuron-specific acid-sensing ion channel 3 by the adaptor protein Na+/H+ exchanger regulatory factor-1. J. Biol. Chem. 281(3), 1796–1807 (2006).

32 Hruska-Hageman AM, Benson CJ, Leonard AS, Price MP, Welsh MJ. PSD-95 and Lin-7b interact with acid-sensing ion channel-3 and have opposite effects on H+- gated current. J. Biol. Chem. 279(45), 46962–46968 (2004).

33 Zha XM, Costa V, Harding AM, Reznikov L, Benson CJ, Welsh MJ. ASIC2 subunits target acid-sensing ion channels to the synapse via an association with PSD-95. J. Neurosci. 29(26), 8438–8446 (2009).

34 Schnizler MK, Schnizler K, Zha XM et al. The cytoskeletal protein alpha-actinin regulates acid-sensing ion channel 1a through a C-terminal interaction. J. Biol. Chem. 284(5), 2697–2705 (2009).

35 Donier E, Rugiero F, Okuse K, Wood JN. Annexin II light chain p11 promotes functional expression of acid-sensing ion channel ASIC1a. J. Biol. Chem. 280(46), 38666–38672 (2005).

36 Price MP, Thompson RJ, Eshcol JO, Wemmie JA, Benson CJ. Stomatin modulates gating of acid-sensing ion channels. J. Biol. Chem. 279(51), 53886–53891 (2004).

37 Wetzel C, Hu J, Riethmacher D et al. A stomatin-domain protein essential for touch sensation in the mouse. Nature 445(7124), 206–209 (2007).

38 Gao J, Duan B, Wang DG et al. Coupling between NMDA receptor and acid-sensing ion channel contributes to ischemic neuronal death. Neuron 48(4), 635–646 (2005).

39 Chai S, Li M, Lan J, Xiong ZG, Saugstad JA, Simon RP. A kinase-anchoring protein 150 and calcineurin are involved in regulation of acid-sensing ion channels ASIC1a and ASIC2a. J. Biol. Chem. 282(31), 22668–22677 (2007).

40 Jasti J, Furukawa H, Gonzales EB, Gouaux E. Structure of acid-sensing ion channel 1 at 1.9 A resolution and low pH. Nature 449(7160), 316–323 (2007).

41 Gonzales EB, Kawate T, Gouaux E. Pore architecture and ion sites in acid-sensing ion channels and P2X receptors. Nature 460(7255), 599–604 (2009).

42 Chen X, Grunder S. Permeating protons contribute to tachyphylaxis of Acid-Sensing Ion Channel (ASIC) 1a. J. Physiol. 579(Pt 3), 657–670 (2007).

43 Sutherland SP, Benson CJ, Adelman JP, McCleskey EW. Acid-sensing ion channel 3 matches the acid-gated current in cardiac ischemia-sensing neurons. Proc. Natl Acad. Sci. USA 98(2), 711–716 (2001).

44 Yermolaieva O, Leonard AS, Schnizler MK, Abboud FM, Welsh MJ. Extracellular acidosis increases neuronal cell calcium by activating acid-sensing ion channel 1a. Proc. Natl Acad. Sci. USA 101(17), 6752–6757 (2004).

45 Samways DS, Harkins AB, Egan TM. Native and recombinant ASIC1a receptors conduct negligible Ca2+ entry. Cell Calcium 45(4), 319–325 (2009).

46 Xiong ZG, Zhu XM, Chu XP et al. Neuroprotection in ischemia: blocking calcium-permeable acid-sensing ion channels. Cell 118(6), 687–698 (2004).

47 Herrera Y, Katnik C, Rodriguez JD et al. sigma-1 receptor modulation of acid-sensing ion channel a (ASIC1a) and ASIC1a-induced Ca2+ influx in rat cortical neurons. J. Pharmacol. Exp. Ther. 327(2), 491–502 (2008).

48 Zha XM, Wemmie JA, Green SH, Welsh MJ. Acid-sensing ion channel 1a is a postsynaptic proton receptor that affects

Page 12: Current perspectives on acid-sensing ion channels: new advances and therapeutic implications

Expert Rev. Clin. Pharmacol. 3(3), (2010)342

Review Noël, Salinas, Baron, Diochot, Deval & Lingueglia

the density of dendritic spines. Proc. Natl Acad. Sci. USA 103(44), 16556–16561 (2006).

49 Donier E, Rugiero F, Jacob C, Wood JN. Regulation of ASIC activity by ASIC4 – new insights into ASIC channel function revealed by a yeast two-hybrid assay. Eur. J. Neurosci. 28(1), 74–86 (2008).

50 Baron A, Voilley N, Lazdunski M, Lingueglia E. Acid sensing ion channels in dorsal spinal cord neurons. J. Neurosci. 28(6), 1498–1508 (2008).

51 Salinas M, Lazdunski M, Lingueglia E. Structural elements for the generation of sustained currents by the acid pain sensor ASIC3. J. Biol. Chem. 284(46), 31851–31859 (2009).

52 Deval E, Baron A, Lingueglia E, Mazarguil H, Zajac JM, Lazdunski M. Effects of neuropeptide SF and related peptides on acid sensing ion channel 3 and sensory neuron excitability. Neuropharmacology 44(5), 662–671 (2003).

53 Yagi J, Wenk HN, Naves LA, McCleskey EW. Sustained currents through ASIC3 ion channels at the modest pH changes that occur during myocardial ischemia. Circ. Res. 99(5), 501–509 (2006).

54 Deval E, Noel J, Lay N et al. ASIC3, a sensor of acidic and primary inflammatory pain. EMBO J. 27(22), 3047–3055 (2008).

55 Li T, Yang Y, Canessa CM. Interaction of the aromatics Tyr-72/Trp-288 in the interface of the extracellular and transmembrane domains is essential for proton gating of acid-sensing ion channels. J. Biol. Chem. 284(7), 4689–4694 (2009).

56 Paukert M, Chen X, Polleichtner G, Schindelin H, Grunder S. Candidate amino acids involved in H+ gating of acid-sensing ion channel 1a. J. Biol. Chem. 283(1), 572–581 (2008).

57 Paukert M, Babini E, Pusch M, Grunder S. Identification of the Ca2+ blocking site of acid-sensing ion channel (ASIC) 1: implications for channel gating. J. Gen. Physiol. 124(4), 383–394 (2004).

58 Yang H, Yu Y, Li WG, Xu TL, Jiang H. Conformational sampling on acid-sensing ion channel 1 (ASIC1): implication for a symmetric conformation. Cell Res. 19(8), 1035–1037 (2009).

59 Yang H, Yu Y, Li WG et al. Inherent dynamics of the acid-sensing ion channel 1 correlates with the gating mechanism. PLoS Biol. 7(7), e1000151 (2009).

60 Immke DC, McCleskey EW. Protons open acid-sensing ion channels by catalyzing relief of Ca2+ blockade. Neuron 37(1), 75–84 (2003).

61 Zhang P, Sigworth FJ, Canessa CM. Gating of acid-sensitive ion channel-1: release of Ca2+ block vs. allosteric mechanism. J. Gen. Physiol. 127(2), 109–117 (2006).

62 Sherwood T, Franke R, Conneely S, Joyner J, Arumugan P, Askwith C. Identification of protein domains that control proton and calcium sensitivity of ASIC1a. J. Biol. Chem. 284(41), 27899–27907 (2009).

63 de Weille J, Bassilana F. Dependence of the acid-sensitive ion channel, ASIC1a, on extracellular Ca(2+) ions. Brain Res. 900(2), 277–281 (2001).

64 Zhang P, Canessa CM. Single channel properties of rat acid-sensitive ion channel-1a, -2a, and -3 expressed in Xenopus oocytes. J. Gen. Physiol. 120(4), 553–566 (2002).

65 Zhang P, Canessa CM. Single-channel properties of recombinant acid-sensitive ion channels formed by the subunits ASIC2 and ASIC3 from dorsal root ganglion neurons expressed in Xenopus oocytes. J. Gen. Physiol. 117(6), 563–572 (2001).

66 Coscoy S, de Weille JR, Lingueglia E, Lazdunski M. The pre-transmembrane 1 domain of acid-sensing ion channels participates in the ion pore. J. Biol. Chem. 274(15), 10129–10132 (1999).

67 Dube GR, Lehto SG, Breese NM et al. Electrophysiological and in vivo characterization of A-317567, a novel blocker of acid sensing ion channels. Pain 117(1–2), 88–96 (2005).

68 Voilley N, de Weille J, Mamet J, Lazdunski M. Nonsteroid anti-inflammatory drugs inhibit both the activity and the inflammation-induced expression of acid-sensing ion channels in nociceptors. J. Neurosci. 21(20), 8026–8033 (2001).

69 Voilley N. Acid-sensing ion channels (ASICs): new targets for the analgesic effects of non-steroid anti-inflammatory drugs (NSAIDs). Curr. Drug Targets Inflamm. Allergy 3(1), 71–79 (2004).

70 Ugawa S, Ishida Y, Ueda T, Inoue K, Nagao M, Shimada S. Nafamostat mesilate reversibly blocks acid-sensing ion channel currents. Biochem. Biophys. Res. Commun. 363(1), 203–208 (2007).

71 Garza A, Lopez-Ramirez O, Vega R, Soto E. The aminoglycosides modulate the acid-sensing ionic-channel (ASIC) currents

in dorsal-root ganglion neurons from the rat. J. Pharmacol. Exp. Ther. 332(2), 489–499 (2009).

72 Chen X, Qiu L, Li M et al. Diarylamidines: High potency inhibitors of acid-sensing ion channels. Neuropharmacology doi: 10.1016/j.neuropharm.2010.01.011 (2010) (Epub ahead of print).

73 Allen NJ, Attwell D. Modulation of ASIC channels in rat cerebellar Purkinje neurons by ischaemia-related signals. J. Physiol. 543(Pt 2), 521–529 (2002).

74 Babinski K, Catarsi S, Biagini G, Seguela P. Mammalian ASIC2a and ASIC3 subunits co-assemble into heteromeric proton-gated channels sensitive to Gd3+. J. Biol. Chem. 275(37), 28519–28525 (2000).

75 Wang W, Yu Y, Xu TL. Modulation of acid-sensing ion channels by Cu(2+) in cultured hypothalamic neurons of the rat. Neuroscience 145(2), 631–641 (2007).

76 Staruschenko A, Dorofeeva NA, Bolshakov KV, Stockand JD. Subunit-dependent cadmium and nickel inhibition of acid-sensing ion channels. Dev. Neurobiol. 67(1), 97–107 (2007).

77 Wang W, Duan B, Xu H, Xu L, Xu TL. Calcium-permeable acid-sensing ion channel is a molecular target of the neurotoxic metal ion lead. J. Biol. Chem. 281(5), 2497–2505 (2006).

78 Chu XP, Wemmie JA, Wang WZ et al. Subunit-dependent high-affinity zinc inhibition of acid-sensing ion channels. J. Neurosci. 24(40), 8678–8689 (2004).

79 Baron A, Schaefer L, Lingueglia E, Champigny G, Lazdunski M. Zn2+ and H+ are coactivators of acid-sensing ion channels. J. Biol. Chem. 276(38), 35361–35367 (2001).

80 Baron A, Waldmann R, Lazdunski M. ASIC-like, proton-activated currents in rat hippocampal neurons. J. Physiol. 539(Pt 2), 485–494 (2002).

81 Diochot S, Salinas M, Baron A, Escoubas P, Lazdunski M. Peptides inhibitors of acid-sensing ion channels. Toxicon 49(2), 271–284 (2007).

82 Escoubas P, De Weille JR, Lecoq A et al. Isolation of a tarantula toxin specific for a class of proton-gated Na+ channels. J. Biol. Chem. 275(33), 25116–25121 (2000).

83 Salinas M, Rash LD, Baron A, Lambeau G, Escoubas P, Lazdunski M. The receptor site of the spider toxin PcTx1 on the proton-gated cation channel ASIC1a. J. Physiol. 570(Pt 2), 339–354 (2006).

Page 13: Current perspectives on acid-sensing ion channels: new advances and therapeutic implications

www.expert-reviews.com 343

ReviewAcid-sensing ion channels: new advances & therapeutic implications

84 Escoubas P, Bernard C, Lambeau G, Lazdunski M, Darbon H. Recombinant production and solution structure of PcTx1, the specific peptide inhibitor of ASIC1a proton-gated cation channels. Protein Sci. 12(7), 1332–1343 (2003).

85 Diochot S, Baron A, Rash LD et al. A new sea anemone peptide, APETx2, inhibits ASIC3, a major acid-sensitive channel in sensory neurons. EMBO J. 23(7), 1516–1525 (2004).

86 Chagot B, Escoubas P, Diochot S, Bernard C, Lazdunski M, Darbon H. Solution structure of APETx2, a specific peptide inhibitor of ASIC3 proton-gated channels. Protein Sci. 14(8), 2003–2010 (2005).

87 Chen X, Kalbacher H, Grunder S. The tarantula toxin psalmotoxin 1 inhibits acid-sensing ion channel (ASIC) 1a by increasing its apparent H+ affinity. J. Gen. Physiol. 126(1), 71–79 (2005).

88 Qadri YJ, Berdiev BK, Song Y, Lippton HL, Fuller CM, Benos DJ. Psalmotoxin-1 docking to human acid-sensing ion channel-1. J. Biol. Chem. 284(26), 17625–17633 (2009).

89 Chen X, Kalbacher H, Grunder S. Interaction of acid-sensing ion channel (ASIC) 1 with the tarantula toxin psalmotoxin 1 is state dependent. J. Gen. Physiol. 127(3), 267–276 (2006).

90 Askwith CC, Cheng C, Ikuma M, Benson C, Price MP, Welsh MJ. Neuropeptide FF and FMRFamide potentiate acid-evoked currents from sensory neurons and proton-gated DEG/ENaC channels. Neuron 26(1), 133–141 (2000).

91 Catarsi S, Babinski K, Seguela P. Selective modulation of heteromeric ASIC proton-gated channels by neuropeptide FF. Neuropharmacology 41(5), 592–600 (2001).

92 Chen X, Paukert M, Kadurin I, Pusch M, Grunder S. Strong modulation by RFamide neuropeptides of the ASIC1b/3 heteromer in competition with extracellular calcium. Neuropharmacology 50(8), 964–974 (2006).

93 Xie J, Price MP, Wemmie JA, Askwith CC, Welsh MJ. ASIC3 and ASIC1 mediate FMRFamide-related peptide enhancement of H+-gated currents in cultured dorsal root ganglion neurons. J. Neurophysiol. 89(5), 2459–2465 (2003).

94 Sherwood TW, Askwith CC. Endogenous arginine-phenylalanine-amide-related peptides alter steady-state desensitization of ASIC1a. J. Biol. Chem. 283(4), 1818–1830 (2008).

95 Askwith CC, Wemmie JA, Price MP, Rokhlina T, Welsh MJ. Acid-sensing ion channel 2 (ASIC2) modulates ASIC1 H+-activated currents in hippocampal neurons. J. Biol. Chem. 279(18), 18296–18305 (2004).

96 Ostrovskaya O, Moroz L, Krishtal O. Modulatory action of RFamide-related peptides on acid-sensing ionic channels is pH dependent: the role of arginine. J. Neurochem. 91(1), 252–255 (2004).

97 Xie J, Price MP, Berger AL, Welsh MJ. DRASIC contributes to pH-gated currents in large dorsal root ganglion sensory neurons by forming heteromultimeric channels. J. Neurophysiol. 87(6), 2835–2843 (2002).

98 Sherwood TW, Askwith CC. Dynorphin opioid peptides enhance acid-sensing ion channel 1a activity and acidosis-induced neuronal death. J. Neurosci. 29(45), 14371–14380 (2009).

99 Wemmie JA, Askwith CC, Lamani E, Cassell MD, Freeman JH, Jr., Welsh MJ. Acid-sensing ion channel 1 is localized in brain regions with high synaptic density and contributes to fear conditioning. J. Neurosci. 23(13), 5496–5502 (2003).

100 Waldmann R, Lazdunski M. H+-gated cation channels: neuronal acid sensors in the NaC/DEG family of ion channels. Curr. Opin. Neurobiol. 8(3), 418–424 (1998).

101 Wemmie JA, Chen J, Askwith CC et al. The acid-activated ion channel ASIC contributes to synaptic plasticity, learning, and memory. Neuron 34(3), 463–477 (2002).

102 Alvarez de la Rosa D, Krueger SR, Kolar A, Shao D, Fitzsimonds RM, Canessa CM. Distribution, subcellular localization and ontogeny of ASIC1 in the mammalian central nervous system. J. Physiol. 546(Pt 1), 77–87 (2003).

103 Roza C, Puel JL, Kress M et al. Knockout of the ASIC2 channel in mice does not impair cutaneous mechanosensation, visceral mechanonociception and hearing. J. Physiol. 558(Pt 2), 659–669 (2004).

104 Cho JH, Askwith CC. Presynaptic release probability is increased in hippocampal neurons from ASIC1 knockout mice. J. Neurophysiol. 99(2), 426–441 (2008).

105 Maren S. An acid-sensing channel sows fear and panic. Cell 139(5), 867–869 (2009).

106 Coryell MW, Ziemann AE, Westmoreland PJ et al. Targeting ASIC1a reduces innate fear and alters neuronal activity in the fear circuit. Biol. Psychiatry 62(10), 1140–1148 (2007).

107 Coryell MW, Wunsch AM, Haenfler JM et al. Restoring acid-sensing ion channel-1a in the amygdala of knock-out mice rescues fear memory but not unconditioned fear responses. J. Neurosci. 28(51), 13738–13741 (2008).

108 Wemmie JA, Coryell MW, Askwith CC et al. Overexpression of acid-sensing ion channel 1a in transgenic mice increases acquired fear-related behavior. Proc. Natl Acad. Sci. USA 101(10), 3621–3626 (2004).

109 Ziemann AE, Allen JE, Dahdaleh NS et al. The amygdala is a chemosensor that detects carbon dioxide and acidosis to elicit fear behavior. Cell 139(5), 1012–1021 (2009).

110 Coryell MW, Wunsch AM, Haenfler JM et al. Acid-sensing ion channel-1a in the amygdala, a novel therapeutic target in depression-related behavior. J. Neurosci. 29(17), 5381–5388 (2009).

111 Friese MA, Craner MJ, Etzensperger R et al. Acid-sensing ion channel-1 contributes to axonal degeneration in autoimmune inflammation of the central nervous system. Nat. Med. 13(12), 1483–1489 (2007).

112 Arias RL, Sung ML, Vasylyev D et al. Amiloride is neuroprotective in an MPTP model of Parkinson’s disease. Neurobiol. Dis. 31(3), 334–341 (2008).

113 Wong HK, Bauer PO, Kurosawa M et al. Blocking acid-sensing ion channel 1 alleviates Huntington’s disease pathology via an ubiquitin-proteasome system-dependent mechanism. Hum. Mol. Genet. 17(20), 3223–3235 (2008).

114 Zhao X, Gorin FA, Berman RF, Lyeth BG. Differential hippocampal protection when blocking intracellular sodium and calcium entry during traumatic brain injury in rats. J. Neurotrauma 25(10), 1195–1205 (2008).

115 Ziemann AE, Schnizler MK, Albert GW et al. Seizure termination by acidosis depends on ASIC1a. Nat. Neurosci. 11(7), 816–822 (2008).

116 Bolshakov KV, Essin KV, Buldakova SL et al. Characterization of acid-sensitive ion channels in freshly isolated rat brain neurons. Neuroscience 110(4), 723–730 (2002).

117 Ali A, Pillai KP, Ahmad FJ, Dua Y, Vohora D. Anticonvulsant effect of amiloride in pentetrazole-induced status epilepticus in mice. Pharmacol. Rep. 58(2), 242–245 (2006).

118 Luszczki JJ, Sawicka KM, Kozinska J, Dudra-Jastrzebska M, Czuczwar SJ. Amiloride enhances the anticonvulsant

Page 14: Current perspectives on acid-sensing ion channels: new advances and therapeutic implications

Expert Rev. Clin. Pharmacol. 3(3), (2010)344

Review Noël, Salinas, Baron, Diochot, Deval & Lingueglia

action of various antiepileptic drugs in the mouse maximal electroshock seizure model. J. Neural. Transm. 116(1), 57–66 (2009).

119 N’Gouemo P. Amiloride delays the onset of pilocarpine-induced seizures in rats. Brain Res. 1222, 230–232 (2008).

120 Miesenbock G, De Angelis DA, Rothman JE. Visualizing secretion and synaptic transmission with pH-sensitive green fluorescent proteins. Nature 394(6689), 192–195 (1998).

121 DeVries SH. Exocytosed protons feedback to suppress the Ca2+ current in mammalian cone photoreceptors. Neuron 32(6), 1107–1117 (2001).

122 Barnes S, Merchant V, Mahmud F. Modulation of transmission gain by protons at the photoreceptor output synapse. Proc. Natl Acad. Sci. USA 90(21), 10081–10085 (1993).

123 Krishtal O. The ASICs: signaling molecules? Modulators? Trends Neurosci. 26(9), 477–483 (2003).

124 Beg AA, Ernstrom GG, Nix P, Davis MW, Jorgensen EM. Protons act as a transmitter for muscle contraction in C. elegans. Cell 132(1), 149–160 (2008).

125 Garcia-Anoveros J, Samad TA, Zuvela-Jelaska L, Woolf CJ, Corey DP. Transport and localization of the DEG/ENaC ion channel BNaC1a to peripheral mechanosensory terminals of dorsal root ganglia neurons. J. Neurosci. 21(8), 2678–2686 (2001).

126 Price MP, McIlwrath SL, Xie J et al. The DRASIC cation channel contributes to the detection of cutaneous touch and acid stimuli in mice. Neuron 32(6), 1071–1083 (2001).

127 Alvarez de la Rosa D, Zhang P, Shao D, White F, Canessa CM. Functional implications of the localization and activity of acid-sensitive channels in rat peripheral nervous system. Proc. Natl Acad. Sci. USA 99(4), 2326–2331 (2002).

128 Price MP, Lewin GR, McIlwrath SL et al. The mammalian sodium channel BNC1 is required for normal touch sensation. Nature 407(6807), 1007–1011 (2000).

129 Molliver DC, Immke DC, Fierro L, Pare M, Rice FL, McCleskey EW. ASIC3, an acid-sensing ion channel, is expressed in metaboreceptive sensory neurons. Mol. Pain 1(1), 35 (2005).

130 Ugawa S, Ueda T, Yamamura H, Shimada S. In situ hybridization evidence for the coexistence of ASIC and TRPV1 within rat single sensory neurons. Brain Res. Mol. Brain Res. 136(1–2), 125–133 (2005).

131 Olson TH, Riedl MS, Vulchanova L, Ortiz-Gonzalez XR, Elde R. An acid sensing ion channel (ASIC) localizes to small primary afferent neurons in rats. Neuroreport 9(6), 1109–1113 (1998).

132 Duan B, Wu LJ, Yu YQ et al. Upregulation of acid-sensing ion channel ASIC1a in spinal dorsal horn neurons contributes to inflammatory pain hypersensitivity. J. Neurosci. 27(41), 11139–11148 (2007).

133 Caterina MJ, Schumacher MA, Tominaga M, Rosen TA, Levine JD, Julius D. The capsaicin receptor: a heat-activated ion channel in the pain pathway. Nature 389(6653), 816–824 (1997).

134 Yudin YK, Tamarova ZA, Ostrovskaya OI, Moroz LL, Krishtal OA. RFa-related peptides are algogenic: evidence in vitro and in vivo. Eur. J. Neurosci. 20(5), 1419–1423 (2004).

135 Mamet J, Baron A, Lazdunski M, Voilley N. Proinflammatory mediators, stimulators of sensory neuron excitability via the expression of acid-sensing ion channels. J. Neurosci. 22(24), 10662–10670 (2002).

136 Chen CC, Zimmer A, Sun WH, Hall J, Brownstein MJ, Zimmer A. A role for ASIC3 in the modulation of high-intensity pain stimuli. Proc. Natl Acad. Sci. USA 99(13), 8992–8997 (2002).

137 Page AJ, Brierley SM, Martin CM et al. The ion channel ASIC1 contributes to visceral but not cutaneous mechanoreceptor function. Gastroenterology 127(6), 1739–1747 (2004).

138 Mogil JS, Breese NM, Witty MF et al. Transgenic expression of a dominant-negative ASIC3 subunit leads to increased sensitivity to mechanical and inflammatory stimuli. J. Neurosci. 25(43), 9893–9901 (2005).

139 Rocha-Gonzalez HI, Herrejon-Abreu EB, Lopez-Santillan FJ, Garcia-Lopez BE, Murbartian J, Granados-Soto V. Acid increases inflammatory pain in rats: effect of local peripheral ASICs inhibitors. Eur. J. Pharmacol. 603(1–3), 56–61 (2009).

140 Yiangou Y, Facer P, Smith JA et al. Increased acid-sensing ion channel ASIC-3 in inflamed human intestine. Eur. J. Gastroenterol. Hepatol. 13(8), 891–896 (2001).

141 Mamet J, Lazdunski M, Voilley N. How nerve growth factor drives physiological and inflammatory expressions of acid-sensing ion channel 3 in sensory neurons. J. Biol. Chem. 278(49), 48907–48913 (2003).

142 Smith ES, Cadiou H, McNaughton PA. Arachidonic acid potentiates acid-sensing ion channels in rat sensory neurons by a direct action. Neuroscience 145(2), 686–698 (2007).

143 Cadiou H, Studer M, Jones NG et al. Modulation of acid-sensing ion channel activity by nitric oxide. J. Neurosci. 27(48), 13251–13260 (2007).

144 Steen KH, Issberner U, Reeh PW. Pain due to experimental acidosis in human skin: evidence for non-adapting nociceptor excitation. Neurosci. Lett. 199(1), 29–32 (1995).

145 Ugawa S, Ueda T, Ishida Y, Nishigaki M, Shibata Y, Shimada S. Amiloride-blockable acid-sensing ion channels are leading acid sensors expressed in human nociceptors. J. Clin. Invest. 110(8), 1185–1190 (2002).

146 Jones NG, Slater R, Cadiou H, McNaughton P, McMahon SB. Acid-induced pain and its modulation in humans. J. Neurosci. 24(48), 10974–10979 (2004).

147 McMahon SB, Jones NG. Plasticity of pain signaling: role of neurotrophic factors exemplified by acid-induced pain. J. Neurobiol. 61(1), 72–87 (2004).

148 Yen YT, Tu PH, Chen CJ, Lin YW, Hsieh ST, Chen CC. Role of acid-sensing ion channel 3 in sub-acute-phase inflammation. Mol. Pain 5, 1 (2009).

149 Mazzuca M, Heurteaux C, Alloui A et al. A tarantula peptide against pain via ASIC1a channels and opioid mechanisms. Nat. Neurosci. 10(8), 943–945 (2007).

150 Leffler A, Monter B, Koltzenburg M. The role of the capsaicin receptor TRPV1 and acid-sensing ion channels (ASICS) in proton sensitivity of subpopulations of primary nociceptive neurons in rats and mice. Neuroscience 139(2), 699–709 (2006).

151 Ikeuchi M, Kolker SJ, Sluka KA. Acid-sensing ion channel 3 expression in mouse knee joint afferents and effects of carrageenan-induced arthritis. J. Pain 10(3), 336–342 (2009).

152 Walder RY, Rasmussen LA, Rainier JD, Light AR, Wemmie JA, Sluka KA. ASIC1 and ASIC3 play different roles in the development of hyperalgesia after inflammatory muscle injury. J. Pain 11(3), 210–218 (2009).

153 Sluka KA, Price MP, Breese NM, Stucky CL, Wemmie JA, Welsh MJ. Chronic hyperalgesia induced by repeated acid injections in muscle is abolished by the loss of ASIC3, but not ASIC1. Pain 106(3), 229–239 (2003).

Page 15: Current perspectives on acid-sensing ion channels: new advances and therapeutic implications

www.expert-reviews.com 345

ReviewAcid-sensing ion channels: new advances & therapeutic implications

154 Sluka KA, Radhakrishnan R, Benson CJ et al. ASIC3 in muscle mediates mechanical, but not heat, hyperalgesia associated with muscle inflammation. Pain 129(1–2), 102–112 (2007).

155 Ikeuchi M, Kolker SJ, Burnes LA, Walder RY, Sluka KA. Role of ASIC3 in the primary and secondary hyperalgesia produced by joint inflammation in mice. Pain 137(3), 662–669 (2008).

156 Burnes LA, Kolker SJ, Danielson JF, Walder RY, Sluka KA. Enhanced muscle fatigue occurs in male but not female ASIC3-/- mice. Am. J. Physiol. Regul. Integr. Comp. Physiol. 294(4), R1347–R1355 (2008).

157 Benson CJ, Sutherland SP. Toward an understanding of the molecules that sense myocardial ischemia. Ann. NY Acad. Sci. 940, 96–109 (2001).

158 Immke DC, McCleskey EW. Lactate enhances the acid-sensing Na+ channel on ischemia-sensing neurons. Nat. Neurosci. 4(9), 869–870 (2001).

159 Sugiura T, Dang K, Lamb K, Bielefeldt K, Gebhart GF. Acid-sensing properties in rat gastric sensory neurons from normal and ulcerated stomach. J. Neurosci. 25(10), 2617–2627 (2005).

160 Bielefeldt K, Davis BM. Differential effects of ASIC3 and TRPV1 deletion on gastroesophageal sensation in mice. Am. J. Physiol. Gastrointest. Liver Physiol. 294(1), G130–138 (2008).

161 Wultsch T, Painsipp E, Shahbazian A et al. Deletion of the acid-sensing ion channel ASIC3 prevents gastritis-induced acid hyperresponsiveness of the stomach-brainstem axis. Pain 134(3), 245–253 (2008).

162 Holzer P. Taste receptors in the gastrointestinal tract. V. Acid sensing in the gastrointestinal tract. Am. J. Physiol. Gastrointest. Liver Physiol. 292(3), G699–G705 (2007).

163 Page AJ, Brierley SM, Martin CM et al. Different contributions of ASIC channels 1a, 2, and 3 in gastrointestinal mechanosensory function. Gut 54(10), 1408–1415 (2005).

164 Hughes PA, Brierley SM, Young RL, Blackshaw LA. Localization and comparative analysis of acid-sensing ion channel (ASIC1, 2, and 3) mRNA expression in mouse colonic sensory neurons within thoracolumbar dorsal root ganglia. J. Comp. Neurol. 500(5), 863–875 (2007).

165 Jones RC 3rd, Xu L, Gebhart GF. The mechanosensitivity of mouse colon afferent fibers and their sensitization by

inflammatory mediators require transient receptor potential vanilloid 1 and acid-sensing ion channel 3. J. Neurosci. 25(47), 10981–10989 (2005).

166 Jones RC, 3rd, Otsuka E, Wagstrom E, Jensen CS, Price MP, Gebhart GF. Short-term sensitization of colon mechanoreceptors is associated with long-term hypersensitivity to colon distention in the mouse. Gastroenterology 133(1), 184–194 (2007).

167 Gu Q, Lee LY. Characterization of acid-signaling in rat vagal pulmonary sensory neurons. Am. J. Physiol. Lung Cell Mol. Physiol. 291(1), L58–L65 (2006).

168 Fukuda T, Ichikawa H, Terayama R, Yamaai T, Kuboki T, Sugimoto T. ASIC3-immunoreactive neurons in the rat vagal and glossopharyngeal sensory ganglia. Brain Res. 1081(1), 150–155 (2006).

169 Groth M, Helbig T, Grau V, Kummer W, Haberberger RV. Spinal afferent neurons projecting to the rat lung and pleura express acid sensitive channels. Respir. Res. 7, 96 (2006).

170 Faisy C, Planquette B, Naline E et al. Acid-induced modulation of airway basal tone and contractility: role of acid-sensing ion channels (ASICs) and TRPV1 receptor. Life Sci. 81(13), 1094–1102 (2007).

171 Canning BJ, Farmer DG, Mori N. Mechanistic studies of acid evoked coughing in anesthetized guinea pigs. Am. J. Physiol. Regul. Integr. Comp. Physiol. 291(2), R454–R463 (2006).

172 Ricciardolo FL. Mechanisms of citric acid-induced bronchoconstriction. Am. J. Med. 111(Suppl. 8A), 18S–24S (2001).

173 Kollarik M, Undem BJ. Mechanisms of acid-induced activation of airway afferent nerve fibres in guinea-pig. J. Physiol. 543(Pt 2), 591–600 (2002).

174 Wu LJ, Duan B, Mei YD et al. Characterization of acid-sensing ion channels in dorsal horn neurons of rat spinal cord. J. Biol. Chem. 279(42), 43716–43724 (2004).

175 Kontinen VK, Aarnisalo AA, Idanpaan-Heikkila JJ, Panula P, Kalso E. Neuropeptide FF in the rat spinal cord during carrageenan inflammation. Peptides 18(2), 287–292 (1997).

176 Yang HY, Iadarola MJ. Activation of spinal neuropeptide FF and the neuropeptide FF receptor 2 during inflammatory hyperalgesia in rats. Neuroscience 118(1), 179–187 (2003).

177 Vilim FS, Aarnisalo AA, Nieminen ML et al. Gene for pain modulatory neuropeptide NPFF: induction in spinal cord by noxious stimuli. Mol. Pharmacol. 55(5), 804–811 (1999).

178 Bounoutas A, Chalfie M. Touch sensitivity in Caenorhabditis elegans. Pflugers Arch. 454(5), 691–702 (2007).

179 O’Hagan R, Chalfie M, Goodman MB. The MEC-4 DEG/ENaC channel of Caenorhabditis elegans touch receptor neurons transduces mechanical signals. Nat. Neurosci. 8(1), 43–50 (2005).

180 Suzuki H, Kerr R, Bianchi L et al. In vivo imaging of C. elegans mechanosensory neurons demonstrates a specific role for the MEC-4 channel in the process of gentle touch sensation. Neuron 39(6), 1005–1017 (2003).

181 Welsh MJ, Price MP, Xie J. Biochemical basis of touch perception: mechanosensory function of degenerin/epithelial Na+ channels. J. Biol. Chem. 277(4), 2369–2372 (2002).

182 Drew LJ, Rohrer DK, Price MP et al. Acid-sensing ion channels ASIC2 and ASIC3 do not contribute to mechanically activated currents in mammalian sensory neurones. J. Physiol. 556(Pt 3), 691–710 (2004).

183 Lu Y, Ma X, Sabharwal R et al. The ion channel ASIC2 is required for baroreceptor and autonomic control of the circulation. Neuron 64(6), 885–897 (2009).

184 Tan ZY, Lu Y, Whiteis CA, Benson CJ, Chapleau MW, Abboud FM. Acid-sensing ion channels contribute to transduction of extracellular acidosis in rat carotid body glomus cells. Circ. Res. 101(10), 1009–1019 (2007).

185 Hayes SG, Kindig AE, Kaufman MP. Blockade of acid sensing ion channels attenuates the exercise pressor reflex in cats. J. Physiol. 581(Pt 3), 1271–1282 (2007).

186 McCord JL, Tsuchimochi H, Kaufman MP. Acid-sensing ion channels contribute to the metaboreceptor component of the exercise pressor reflex. Am. J. Physiol. Heart Circ. Physiol. 297(1), H443–H449 (2009).

187 Lilley S, LeTissier P, Robbins J. The discovery and characterization of a proton-gated sodium current in rat retinal ganglion cells. J. Neurosci. 24(5), 1013–1022 (2004).

188 Ettaiche M, Guy N, Hofman P, Lazdunski M, Waldmann R. Acid-sensing ion channel 2 is important for retinal function and

Page 16: Current perspectives on acid-sensing ion channels: new advances and therapeutic implications

Expert Rev. Clin. Pharmacol. 3(3), (2010)346

Review Noël, Salinas, Baron, Diochot, Deval & Lingueglia

protects against light-induced retinal degeneration. J. Neurosci. 24(5), 1005–1012 (2004).

189 Brockway LM, Zhou ZH, Bubien JK, Jovov B, Benos DJ, Keyser KT. Rabbit retinal neurons and glia express a variety of ENaC/DEG subunits. Am. J. Physiol. Cell Physiol. 283(1), C126–C134 (2002).

190 Ettaiche M, Deval E, Cougnon M, Lazdunski M, Voilley N. Silencing acid-sensing ion channel 1a alters cone-mediated retinal function. J. Neurosci. 26(21), 5800–5809 (2006).

191 Ettaiche M, Deval E, Pagnotta S, Lazdunski M, Lingueglia E. Acid-sensing ion channel 3 in retinal function and survival. Invest. Ophthalmol. Vis. Sci. 50(5), 2417–2426 (2009).

192 Hildebrand MS, de Silva MG, Klockars T et al. Characterisation of DRASIC in the mouse inner ear. Heart Res. 190(1–2), 149–160 (2004).

193 Peng BG, Ahmad S, Chen S, Chen P, Price MP, Lin X. 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(45), 10167–10175 (2004).

194 Ugawa S, Inagaki A, Yamamura H et al. Acid-sensing ion channel-1b in the stereocilia of mammalian cochlear hair cells. Neuroreport 17(12), 1235–1239 (2006).

195 Zhang M, Gong N, Lu YG, Jia NL, Xu TL, Chen L. Functional characterization of acid-sensing ion channels in cultured neurons of rat inferior colliculus. Neuroscience 154(2), 461–472 (2008).

196 Wu WL, Wang CH, Huang EY, Chen CC. Asic3(-/-) female mice with hearing deficit affects social development of pups. PLoS One 4(8), e6508 (2009).

197 Ugawa S, Minami Y, Guo W et al. Receptor that leaves a sour taste in the mouth. Nature 395(6702), 555–556 (1998).

198 Ugawa S, Yamamoto T, Ueda T et al. Amiloride-insensitive currents of the acid-sensing ion channel-2a (ASIC2a)/ASIC2b heteromeric sour-taste receptor channel. J. Neurosci. 23(9), 3616–3622 (2003).

199 Richter TA, Dvoryanchikov GA, Roper SD, Chaudhari N. Acid-sensing ion channel-2 is not necessary for sour taste in mice. J. Neurosci. 24(16), 4088–4091 (2004).

200 Lin W, Ogura T, Kinnamon SC. Acid-activated cation currents in rat vallate taste receptor cells. J. Neurophysiol. 88(1), 133–141 (2002).

201 Liu L, Simon SA. Acidic stimuli activates two distinct pathways in taste receptor cells from rat fungiform papillae. Brain Res. 923(1–2), 58–70 (2001).

202 Huque T, Cowart BJ, Dankulich-Nagrudny L et al. Sour ageusia in two individuals implicates ion channels of the ASIC and PKD families in human sour taste perception at the anterior tongue. PLoS One 4(10), e7347 (2009).

203 Su X, Li Q, Shrestha K et al. Interregulation of proton-gated Na+ channel 3 and cystic fibrosis transmembrane conductance regulator. J. Biol. Chem. 281(48), 36960–36968 (2006).

204 Agopyan N, Bhatti T, Yu S, Simon SA. Vanilloid receptor activation by 2- and 10-microm particles induces responses leading to apoptosis in human airway epithelial cells. Toxicol. Appl. Pharmacol. 192(1), 21–35 (2003).

205 Kullmann FA, Shah MA, Birder LA, de Groat WC. Functional TRP and ASIC-like channels in cultured urothelial cells from the rat. Am. J. Physiol. Renal Physiol. 296(4), F892–F901 (2009).

206 Huang SJ, Yang WS, Lin YW, Wang HC, Chen CC. Increase of insulin sensitivity and reversal of age-dependent glucose intolerance with inhibition of ASIC3. Biochem. Biophys. Res. Commun. 371(4), 729–734 (2008).

207 Jahr H, van Driel M, van Osch GJ, Weinans H, van Leeuwen JP. Identification of acid-sensing ion channels in bone. Biochem. Biophys. Res. Commun. 337(1), 349–354 (2005).

208 Uchiyama Y, Cheng CC, Danielson KG et al. Expression of acid-sensing ion channel 3 (ASIC3) in nucleus pulposus cells of the intervertebral disc is regulated by p75NTR and ERK signaling. J. Bone Miner. Res. 22(12), 1996–2006 (2007).

209 Kolker SJ, Walder RY, Usachev Y et al. ASIC3 expressed in Type B synoviocytes and chondrocytes modulates hyaluronan expression and release. Ann. Rheum. Dis. DOI: 10.1136/ard.2009.117168 (2009) (Epub ahead of print).

210 Drummond HA, Grifoni SC, Jernigan NL. A new trick for an old dogma: ENaC proteins as mechanotransducers in vascular smooth muscle. Physiology (Bethesda) 23, 23–31 (2008).

211 Grifoni SC, Jernigan NL, Hamilton G, Drummond HA. ASIC proteins regulate smooth muscle cell migration. Microvasc. Res. 75(2), 202–210 (2008).

212 Grifoni SD, McKey SE, Drummond HA. Hsc70 regulates cell surface ASIC2 expression and vascular smooth muscle cell migration. Am. J. Physiol. Heart Circ. Physiol. 294(5), H2022–H2030 (2008).

213 Gannon KP, Vanlandingham LG, Jernigan NL, Grifoni SC, Hamilton G, Drummond HA. Impaired pressure-induced constriction in mouse middle cerebral arteries of ASIC2 knockout mice. Am. J. Physiol. Heart Circ. Physiol. 294(4), H1793–H1803 (2008).

214 Vila-Carriles WH, Kovacs GG, Jovov B et al. Surface expression of ASIC2 inhibits the amiloride-sensitive current and migration of glioma cells. J. Biol. Chem. 281(28), 19220–19232 (2006).

215 Bernardinelli L, Murgia SB, Bitti PP et al. Association between the ACCN1 gene and multiple sclerosis in Central East Sardinia. PLoS One 2(5), e480 (2007).

216 Girirajan S, Williams SR, Garbern JY, Nowak N, Hatchwell E, Elsea SH. 17p11.2p12 triplication and del(17)q11.2q12 in a severely affected child with dup(17)p11.2p12 syndrome. Clin. Genet. 72(1), 47–58 (2007).

217 Stone JL, Merriman B, Cantor RM, Geschwind DH, Nelson SF. High density SNP association study of a major autism linkage region on chromosome 17. Hum. Mol. Genet. 16(6), 704–715 (2007).

218 Hettema JM, An SS, Neale MC, van den Oord EJ, Kendler KS, Chen X. Lack of association between the amiloride-sensitive cation channel 2 (ACCN2) gene and anxiety spectrum disorders. Psychiatr. Genet. 18(2), 73–79 (2008).