20.11.2014 Views

Differential maturation of chloride homeostasis

Differential maturation of chloride homeostasis

Differential maturation of chloride homeostasis

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

This article was published in an Elsevier journal. The attached copy<br />

is furnished to the author for non-commercial research and<br />

education use, including for instruction at the author’s institution,<br />

sharing with colleagues and providing to institution administration.<br />

Other uses, including reproduction and distribution, or selling or<br />

licensing copies, or posting to personal, institutional or third party<br />

websites are prohibited.<br />

In most cases authors are permitted to post their version <strong>of</strong> the<br />

article (e.g. in Word or Tex form) to their personal website or<br />

institutional repository. Authors requiring further information<br />

regarding Elsevier’s archiving and manuscript policies are<br />

encouraged to visit:<br />

http://www.elsevier.com/copyright


Author's personal copy<br />

Int. J. Devl Neuroscience 25 (2007) 479–489<br />

www.elsevier.com/locate/ijdevneu<br />

<strong>Differential</strong> <strong>maturation</strong> <strong>of</strong> <strong>chloride</strong> <strong>homeostasis</strong> in primary<br />

afferent neurons <strong>of</strong> the somatosensory system<br />

Daniel Gilbert a,1 , Christina Franjic-Würtz a , Katharina Funk a ,<br />

Thomas Gensch b , Stephan Frings a, *, Frank Möhrlen a<br />

a Department <strong>of</strong> Molecular Physiology, University <strong>of</strong> Heidelberg, Im Neuenheimer Feld 230, 69120 Heidelberg, Germany<br />

b Institute for Neuroscience and Biophysics 1, Juelich Research Center, Leo-Brand-Strasse, 52425 Juelich, Germany<br />

Received 16 June 2007; received in revised form 23 July 2007; accepted 6 August 2007<br />

Abstract<br />

Recent research into the generation <strong>of</strong> hyperalgesia has revealed that both the excitability <strong>of</strong> peripheral nociceptors and the transmission <strong>of</strong> their<br />

afferent signals in the spinal cord are subject to modulation by Cl currents. The underlying Cl <strong>homeostasis</strong> <strong>of</strong> nociceptive neurons, in particular<br />

its postnatal <strong>maturation</strong>, is, however, poorly understood. Here we measure the intracellular Cl concentration, [Cl ] i , <strong>of</strong> somatosensory neurons in<br />

intact dorsal root ganglia <strong>of</strong> mice. Using two-photon fluorescence-lifetime imaging microscopy, we determined [Cl ] i in newborn and adult<br />

animals. We found that the somatosensory neurons undergo a transition <strong>of</strong> Cl <strong>homeostasis</strong> during the first three postnatal weeks that leads to a<br />

decline <strong>of</strong> [Cl ] i in most neurons. Immunohistochemistry showed that a major fraction <strong>of</strong> neurons in the dorsal root ganglia express the cation–<br />

<strong>chloride</strong> co-transporters NKCC1 and KCC2, indicating that the molecular equipment for Cl accumulation and extrusion is present. RT-PCR<br />

analysis showed that the transcription pattern <strong>of</strong> electroneutral Cl co-transporters does not change during the <strong>maturation</strong> process.<br />

These findings demonstrate that dorsal root ganglion neurons undergo a developmental transition <strong>of</strong> <strong>chloride</strong> <strong>homeostasis</strong> during the first three<br />

postnatal weeks. This process parallels the developmental ‘‘<strong>chloride</strong> switch’’ in the central nervous system. However, while most CNS neurons<br />

achieve homogeneously low [Cl ] i levels – which is the basis <strong>of</strong> GABAergic and glycinergic inhibition – somatosensory neurons maintain a<br />

heterogeneous pattern <strong>of</strong> [Cl ] i values. This suggests that Cl currents are excitatory in some somatosensory neurons, but inhibitory in others. Our<br />

results are consistent with the hypothesis that Cl <strong>homeostasis</strong> in somatosensory neurons is regulated through posttranslational modification <strong>of</strong><br />

cation–<strong>chloride</strong> co-transporters.<br />

# 2007 ISDN. Published by Elsevier Ltd. All rights reserved.<br />

Keywords: Pain; Nociceptors; Dorsal root ganglia; Chloride <strong>homeostasis</strong>; Chloride transporters; Fluorescence-lifetime imaging<br />

1. Introduction<br />

Neuronal activity can be strongly influenced by Cl currents.<br />

Inhibitory Cl currents mediate GABAergic and glycinergic<br />

inhibition in the CNS. Excitatory Cl currents occur in neurons<br />

which accumulate Cl , as described for immature neurons <strong>of</strong> the<br />

CNS (Ben-Ari, 2002), for olfactory sensory neurons (Kaneko<br />

Abbreviations: 2P-FLIM, two-photon fluorescence-lifetime imaging<br />

microscopy; [Cl ] i , intracellular <strong>chloride</strong> concentration; CGRP, calcitoningene<br />

related peptide; MQAE, N-6-methoxyquinolinium acetoethylester;<br />

TRPV1, transient receptor potential vanilloid receptor type 1.<br />

* Corresponding author. Tel.: +49 6221 54 5661; fax: +49 6221 54 5627.<br />

E-mail address: s.frings@zoo.uni-heidelberg.de (S. Frings).<br />

1 Current address: Department <strong>of</strong> Physiology and Pharmacology, University<br />

<strong>of</strong> Queensland, Brisbane, QLD 4072, Australia.<br />

et al., 2004), for neurons in spinal and autonomic ganglia (review,<br />

Frings et al., 2000), as well as for neurons challenged by<br />

ischemia, inflammation, or neurological disorders (Payne et al.,<br />

2003; Pond et al., 2006; De Koninck, 2007). The balance<br />

between inhibitory and excitatory Cl effects is determined by<br />

Cl uptake and Cl extrusion pathways active in each<br />

cell. Studies <strong>of</strong> neuronal Cl <strong>homeostasis</strong> have indicated that<br />

the Na + –K + –2Cl co-transporter NKCC1 provides the main<br />

route for Cl uptake, while KCC2 couples Cl extrusion to K +<br />

efflux. It is generally held that NKCC1 expression dominates in<br />

immature neurons <strong>of</strong> cortex and hippocampus, and that KCC2<br />

expression is only induced after birth (Lu et al., 1999; Rivera<br />

et al., 1999; Stein et al., 2004). This developmental transition –<br />

<strong>of</strong>ten termed ‘‘<strong>chloride</strong> switch’’ – shifts the Cl -equilibrium<br />

potential, E Cl , from values near 40 mV to values near the<br />

resting voltage. Once this transition is completed (in rats and<br />

0736-5748/$30.00 # 2007 ISDN. Published by Elsevier Ltd. All rights reserved.<br />

doi:10.1016/j.ijdevneu.2007.08.001


Author's personal copy<br />

480<br />

D. Gilbert et al. / Int. J. Devl Neuroscience 25 (2007) 479–489<br />

mice 2 weeks after birth), the opening <strong>of</strong> Cl channels stabilizes<br />

the resting voltage.<br />

In contrast to CNS neurons, primary afferent neurons <strong>of</strong> the<br />

peripheral nervous system (e.g. somatosensory neurons and<br />

olfactory sensory neurons) are believed to mature without<br />

undergoing a transition <strong>of</strong> <strong>chloride</strong> <strong>homeostasis</strong>, and to support<br />

depolarizing Cl currents throughout the adult life. In olfactory<br />

sensory neurons, E Cl is maintained near 0 mV in adult animals,<br />

and a depolarizing Cl efflux carries the main part <strong>of</strong> the odorinduced<br />

receptor current (Reuter et al., 1998; Kaneko et al.,<br />

2004). A similar role is assumed for Cl currents in<br />

somatosensory neurons, based on reports <strong>of</strong> elevated Cl<br />

levels (30–50 mM) in dorsal root ganglion neurons (Gallagher<br />

et al., 1978; Alvarez-Leefmans et al., 1988; Kenyon, 2001;<br />

Kaneko et al., 2002). Furthermore, somatosensory neurons can<br />

respond to GABA with depolarization (Rudomin and Schmidt,<br />

1999; Sung et al., 2000). Finally, the expression <strong>of</strong> NKCC1<br />

(Alvarez-Leefmans et al., 2001; Kanaka et al., 2001), combined<br />

with the hardly detectable levels <strong>of</strong> KCC2 mRNA (Rivera et al.,<br />

1999; Kanaka et al., 2001), is interpreted as indicative <strong>of</strong> Cl<br />

accumulation. However, two observations are not easily<br />

reconciled with the general concept <strong>of</strong> excitatory Cl currents<br />

in somatosensory neurons: (1) the measured E Cl values (around<br />

40 mV) limit the excitatory effect <strong>of</strong> Cl currents and may<br />

even oppose strong depolarization and excitation; (2) NKCC1<br />

mRNA was detected in only 50% <strong>of</strong> somatosensory neurons in<br />

dorsal root ganglia and trigeminal ganglia, where it is colocalized<br />

with TRPV1 (Price et al., 2006). This expression<br />

pattern may indicate that excitatory Cl currents are restricted<br />

to certain modalities <strong>of</strong> the somatosensory system.<br />

To approach the question <strong>of</strong> Cl accumulation and its role in<br />

the somatosensory system, we examined the <strong>maturation</strong> <strong>of</strong> Cl<br />

<strong>homeostasis</strong> after birth. We directly measured [Cl ] i in intact<br />

dorsal root ganglion neurons <strong>of</strong> early postnatal and adult mice<br />

using two-photon fluorescence-lifetime imaging microscopy<br />

(2P-FLIM). We looked at the expression <strong>of</strong> NKCC1 and KCC2<br />

protein in rat dorsal root ganglia. To assess possible<br />

contributions <strong>of</strong> other Cl transporters to the <strong>maturation</strong> <strong>of</strong><br />

Cl <strong>homeostasis</strong>, we examined the transcription pattern <strong>of</strong> all<br />

genes that encode electroneutral cation–<strong>chloride</strong> co-transporters<br />

in newborn and adult rats.<br />

2. Experimental procedures<br />

2.1. Isolation <strong>of</strong> somatosensory neurons for calibration<br />

All experimental procedures were performed in accordance with the Animal<br />

Protection Law and the guidelines and permissions <strong>of</strong> Heidelberg University.<br />

Adult female NMRI (Naval Medical Research Institute, USA) mice were<br />

anesthetized and killed by cervical transsection. The spinal column was<br />

prepared and opened sagittally with scissors. Thirty to forty dorsal root ganglia<br />

were dissected from the lumbar, thoracic and cervical region. Ganglia were cut<br />

into small pieces and incubated in 2 ml <strong>of</strong> collagenase solution (0.3% collagenase,<br />

C-9891, Sigma, Germany, in DMEM) for 1 h at 37 8C. After<br />

centrifugation (200 g, 20 min) the supernatant was discarded, and 2 ml <strong>of</strong><br />

trypsin solution (0.25% trypsin, T-1426, Sigma, in MEM) was added to the<br />

pellet. The pellet was resuspended by trituration with a truncated, fire-polished<br />

Pasteur pipette (3 mm opening). The ganglia were kept in trypsin solution for<br />

30 min at 37 8C, and then centrifuged for 20 min at 200 g. The supernatant<br />

was discarded and 4 ml <strong>of</strong> culture medium (DMEM + 10% FCS, 10106-169,<br />

and 1% antibiotics/antimycotics, 15240-062, Gibco Life Technologies, Invitrogen,<br />

Karlsruhe, Germany) was added to the pellet. The pellet was triturated, and<br />

then filtered through a 150 mm nylon mesh (Typ 1110, Bückmann, Germany).<br />

This step removes a substantial amount <strong>of</strong> myelin debris and non-dissociated<br />

fragments <strong>of</strong> ganglia, which are retained on the nylon mesh. The pooled cells<br />

were plated on coverslips (coated with 100 mg/ml poly-L-lysine, P-1399,<br />

Sigma, and 20 mg/ml laminin, L-2020, Sigma) and cultured at 37 8C in culture<br />

medium under an atmosphere containing 5% CO 2 . Nerve growth factor (NGF-b<br />

from rat, N-2513, Sigma) was added 1 h after plating at a final concentration <strong>of</strong><br />

50 ng/ml. Cultures were used within 24 h after plating. Cells were incubated in<br />

standard extracellular solution (Table 1) containing 5 mM MQAE (N-6-methoxyquinolinium<br />

acetoethylester; Molecular Probes, Invitrogen) for at least<br />

30 min in the dark, at 37 8C. The dye progressively accumulates within the<br />

cells as the molecule is rendered membrane impermeable by cytosolic esterases<br />

(Koncz and Daugirdas, 1994) with only minor effects on its fluorescence<br />

properties (Kaneko et al., 2002). After incubation, extracellular MQAE was<br />

washed away and coverslips were transferred to the FLIM instrument.<br />

2.2. Tissue preparations for 2P-FLIM measurements<br />

Newborn or adult female NMRI mice were anesthetized and killed by<br />

cervical transsection. The backbone was prepared and vertebrae containing the<br />

spinal cord with attached dorsal root ganglia were isolated with a razor blade<br />

from the spinal column. All preparations were done in cooled (4 8C) standard<br />

extracellular solution (Table 1). Only those vertebrae were used for 2P-FLIM<br />

measurements whose dorsal root ganglia had an intact dura mater and which<br />

were connected to undamaged dorsal roots and spinal nerves. Preparations were<br />

incubated in standard extracellular solution containing 5 mM MQAE for 3.5 h<br />

at room temperature (20 8C). After incubation, the dye solution was replaced by<br />

standard extracellular solution and the preparations were stored at 4 8C until the<br />

measurement. For 2P-FLIM recordings, MQAE-loaded preparations were<br />

immobilized with low melting agarose (USB Corporation, USA) at the bottom<br />

<strong>of</strong> 6 cm dishes (Greiner, Germany). All recordings were done at room temperature<br />

in standard extracellular solution. To test the viability <strong>of</strong> neurons in the<br />

preparation during the time <strong>of</strong> the 2P-FLIM experiments, we performed<br />

viability tests with 0.5 mg/ml MTT (3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium<br />

bromide; Molecular Probes), and with 0.4% trypan blue (Gibco<br />

Life Technologies). Following the preparation, the tissues were incubated at<br />

room temperature in standard extracellular solution and were subsequently<br />

loaded with either MTT (60 min) or trypan blue (10 min) at various times.<br />

Using this method, we established a time window <strong>of</strong> 5 h post mortem during<br />

which the viability <strong>of</strong> DRG neurons was good and the tissue could be used for<br />

2P-FLIM analysis.<br />

2.3. Two-photon fluorescence-lifetime imaging microscopy (2P-<br />

FLIM)<br />

MQAE was used as a fluorescent probe for intracellular Cl ([Cl ] i )<br />

(Verkman, 1990). MQAE molecules reach the excited state upon absorption<br />

<strong>of</strong> a single ultraviolet photon (l = 375 nm) or, alternatively, the simultaneous<br />

absorption <strong>of</strong> two infrared photons (l = 750 nm). We used two-photon excita-<br />

Table 1<br />

Solutions (in mM)<br />

Na K Ca Mg Cl NO 3<br />

Standard extracellular solution 140 5 2.5 1 151<br />

15 Cl standard solution 150 15 105<br />

30 Cl standard solution 150 30 120<br />

45 Cl standard solution 150 45 105<br />

60 Cl standard solution 150 60 90<br />

75 Cl standard solution 150 75 75<br />

Solutions contained 10 mM glucose; pH 7.4 was buffered with 10 mM HEPES.<br />

The cell isolation solution was buffered with phosphate (in mM: 1.9 NaH 2 PO 4 ,<br />

8.1 Na 2 HPO 4 ).


Author's personal copy<br />

D. Gilbert et al. / Int. J. Devl Neuroscience 25 (2007) 479–489 481<br />

tion to achieve an optical resolution <strong>of</strong> 0.5 mm(x-/y-plane) and 1 mm(z-axis).<br />

The infrared light used for two-photon excitation caused no detectable photodamage,<br />

even with the relatively long observation times <strong>of</strong> multiple images with<br />

1 min <strong>of</strong> illumination per image. For the MQAE molecule, the dwell time in the<br />

excited singlet state (the fluorescence lifetime, t) is near 30 ns in water<br />

containing 50 mM MQAE and is reduced by anions through collisional quenching.<br />

The Cl dependence <strong>of</strong> t is described by the Stern–Volmer relation<br />

(t 0 /t =1+K SV [Cl ] i ), where t 0 is the fluorescence lifetime in 0 mM Cl<br />

and K SV , the Stern–Volmer constant, is a measure <strong>of</strong> the Cl sensitivity <strong>of</strong><br />

MQAE. K SV has a value <strong>of</strong> 185 M 1 in water but only 3–20 M 1 inside cells<br />

(Lau et al., 1994; Bevensee et al., 1997; Maglova et al., 1998; Eberhardson et al.,<br />

2000; Kaneko et al., 2002). This reduced sensitivity <strong>of</strong> intracellular MQAE may<br />

result in part from interactions <strong>of</strong> the dye with other soluble anions and from<br />

self-quenching <strong>of</strong> MQAE at concentrations >100 mM(Kaneko et al., 2002). For<br />

2P-FLIM measurements, the tissues or cells were placed on the stage <strong>of</strong> an<br />

upright fluorescence microscope (BX50WI; Olympus Optical, Japan) and<br />

observed through a 60 water-immersion objective (n.a. = 0.9; Olympus<br />

Optical). Fluorescence was excited with 150 fs light pulses (l = 750 nm)<br />

applied at sufficient intensity to generate two-photon excitation. Light pulses<br />

were generated at a frequency <strong>of</strong> 75 MHz by a mode-locked Titan-Sapphire<br />

laser (Mira 900; output power >500 mW; Coherent, Santa Clara, USA), which<br />

was pumped by the frequency-doubled output (532 nm) <strong>of</strong> a Nd–vanadate laser<br />

(Verdi; Coherent). The laser light was directed through the objective to the<br />

tissue surface at reduced power (2.5 mW) using a beam scanner (TILL<br />

Photonics, Munich, Germany). Fluorescence was recorded by photomultipliers,<br />

and lifetime analysis was performed using electronics (SPC-730; Becker &<br />

Hickl, Berlin, Germany) and s<strong>of</strong>tware (SPC7.22; Becker & Hickl) for timecorrelated<br />

single-photon counting (Lakowicz, 1999). Lifetime images were<br />

analyzed using SPCImage 1.8 and 2.6 (Becker & Hickl) and a self-made image<br />

analysis s<strong>of</strong>tware. A detailed description <strong>of</strong> the instrument and the calibration<br />

procedure is described in Kaneko et al. (2002). Images were obtained by<br />

scanning the excitation light focus through tissue layers 1–3 cells below the<br />

dura mater. Mean values <strong>of</strong> Cl concentrations are given with standard<br />

deviations (S.D.).<br />

2.4. Single cell-based quantitative analysis <strong>of</strong> 2P-FLIM data<br />

For quantification <strong>of</strong> cytosolic fluorescence lifetimes, we developed a<br />

s<strong>of</strong>tware that enables single cell-based analysis in 2P-FLIM z-scan images.<br />

The cytosolic MQAE signal <strong>of</strong> a single neuron was analyzed at the equatorial<br />

layer, i.e. the optical section in the image stack, where the total area <strong>of</strong> the cell<br />

was largest. The cytosolic region was outlined by defining manually two regions<br />

<strong>of</strong> interest, delimiting the margins <strong>of</strong> the nucleus and <strong>of</strong> the cell, respectively.<br />

The quantitative parameters, mean fluorescence lifetime (in ns) and cell<br />

diameter (in mm), were extracted by the s<strong>of</strong>tware. The cell diameter, d, was<br />

calculated according to d = H(4A/p), with A = area <strong>of</strong> the cell, assuming<br />

circular cell morphology. The diameter is an important parameter, because<br />

sensory modalities <strong>of</strong> dorsal root ganglion neurons are <strong>of</strong>ten correlated with this<br />

value (e.g. Petruska et al., 2002).<br />

2.5. Immunohistochemistry<br />

To examine the expression <strong>of</strong> Cl transport proteins, we used rat instead <strong>of</strong><br />

mouse dorsal root ganglia because the available antibodies were better suited<br />

for rat than for mouse tissue. Four dorsal root ganglia were dissected from the<br />

thoracic spinal cord <strong>of</strong> two adult Wistar rats and immersed in 4% paraformaldehyde<br />

in PBS (8.1 mM Na 2 HPO 4 , 1.9 mM NaH 2 PO 4 , 130 mM NaCl, pH 7.4)<br />

for 25 min. After three washing steps in PBS, the tissue was sequentially<br />

dehydrated in 10–30% sucrose overnight. Tissue was then embedded in Tissue-<br />

Tek (Leica, Nussloch, Germany) and frozen onto the cryostate stage (Leica CM<br />

3050 S). Four to eight cryosections (12 mm) were cut and collected on coated<br />

slides (SuperFrost 1 Plus, Menzel, Braunschweig, Germany). Sections were airdried<br />

for 30 min, post-fixed in 4% paraformaldehyde in PBS, washed three<br />

times in PBS and blocked in 5% ChemiBLOCKER TM (Millipore, Billerica,<br />

USA), 0.5% Triton X100 in PBS. Primary antibodies were diluted in the same<br />

buffer and incubated for 2 h. The following antibodies and dilutions were used:<br />

mouse anti-NKCC1 1:20 (T4, monoclonal antibody, obtained from the Developmental<br />

Studies Hybridoma Bank, The University <strong>of</strong> Iowa), goat anti-NKCC1<br />

1:20 (N-16; Santa Cruz Biotech, Heidelberg, Germany, #sc-21545), goat anti-<br />

KCC2 1:20 (Santa Cruz Biotech, #sc-19420). After three washing steps in PBS,<br />

sections were incubated for 90 min in fluorescent-labeled donkey anti-mouse<br />

Alexa594 (Invitrogen, A-21203, 1:500) or donkey anti-goat Alexa488 (Invitrogen,<br />

A-21206, 1:500) secondary antibodies. After three washing steps in PBS, a<br />

0.3 mM DAPI solution (Fluka, #32670) was used to stain the nuclei. Sections<br />

were coverslipped with Aqua Poly/Mount (Polysciences, Warrington, USA) and<br />

analyzed using a Nikon Eclipse 90i upright automated microscope equipped<br />

with a Nikon DS-1QM CCD camera. The instrument was used at the Nikon<br />

Imaging Center at the University <strong>of</strong> Heidelberg. No fluorescence signal was<br />

observed upon omission <strong>of</strong> primary antibodies. To test for the specificity <strong>of</strong><br />

primary antibodies, preadsorption controls were performed for the NKCC1 (N-<br />

16) and the KCC2 polyclonal antibodies using a fivefold excess <strong>of</strong> blocking<br />

peptides (0.05 mg/ml) according to the manufacturer’s specifications. No<br />

immunosignal was observed after preadsorption. The specificity <strong>of</strong> the monoclonal<br />

T4 antibody for NKCC1 was demonstrated by Chen et al. (2005) through<br />

the absence <strong>of</strong> T4 immunosignals in brain tissue <strong>of</strong> NKCC1 / mice. Two slices<br />

were evaluated for each test. To avoid double counting <strong>of</strong> cells, every fourth to<br />

sixth section was evaluated, ensuring a minimum distance <strong>of</strong> 24–48 mm<br />

between individual sections.<br />

2.6. Expression <strong>of</strong> electroneutral cation–<strong>chloride</strong> co-transporter<br />

genes<br />

Dorsal root ganglia and, for controls, kidney and brain were dissected from<br />

adult and newborn (P1) Wistar rats. Total RNA was extracted following the<br />

protocol <strong>of</strong> Chomczynski and Sacchi (1987). After DNase I treatment (RNasefree,<br />

Fermentas, StLeon-Rot, Germany) cDNA was synthesized from 5 mg total<br />

RNA using an oligo-dT primer and Superscript TM III Reverse Transcriptase<br />

(Invitrogen) according to the manufacturer’s instructions. The cDNA was<br />

quantified for normalization using PCR (16, 20, 24, 28 and 32 cycles) with<br />

b-actin and ATPase primers. The annealing temperature for all primers was<br />

58 8C. Semi-quantitative PCR amplification was performed on 0.5 ml singlestranded<br />

cDNA product with 2U Taq DNA polymerase (Fermentas). The<br />

cycling conditions were 94 8C for 3 min, 94 8C for 30 s, 58 8C for 30 s,<br />

72 8C for 1 min for 28 and 32 cycles, respectively, and 72 8C for 8 min. For<br />

each cation–<strong>chloride</strong> co-transporter gene, primer pairs were: CCC9/F, 5 0 -TCA<br />

CTG TGT TTG GGG TGT TC-3 0 ; CCC9/R, 5 0 -GGA GAG GGC GAA GTA<br />

AGA GTA G-3 0 ; CCC6/F, 5 0 -GGT TCA ACG GAA GCA CCC TAA-3 0 ; CCC6/<br />

R, 5 0 -GTG ACC ACA GCA GCC AAT GT-3 0 ; KCC1/F, 5 0 -TGA CCC TAG<br />

TGG TGT TTG TCG G-3 0 ; KCC1/R, 5 0 -GTT CCT GCT GAC GCC ATC A-3 0 ;<br />

KCC2/F, 5 0 -GTC TCT GGG CCC GGA GTT T-3 0 ; KCC2/R, 5 0 -GGC ATC CCG<br />

CAG GTC TC-3 0 ; KCC3/F, 5 0 -TGC TGC TGT ACA ATG TTA ACT GCC-3 0 ;<br />

KCC3/R, 5 0 -TAG CCA ACC CTG GAATGC C-3 0 ; KCC4/F, 5 0 -TGC TTT CTA<br />

TCC TGG CCA TCT ATG-3 0 ; KCC4/R, 5 0 -GCC TCC CCA AAC TTA TCT<br />

CGC-3 0 ; NKCC1/F, 5 0 -CGA ATTATT GGA GCC ATTACA GT-3 0 ; NKCC1/R,<br />

5 0 -ACATCT GGA AAG CTG GGT AGATA-3 0 ; NKCC2/F, 5 0 -ATT CAATGA<br />

TGG TGG ATC CAA C-3 0 ; NKCC2/R, 5 0 -CGG CGATGA GAATGA ATG C-<br />

3 0 ; TSC/F, 5 0 -GTA GAC CCC ATC AAT GAC ATC C-3 0 ; TSC/R, 5 0 -AAG CCA<br />

ATC AGA GGG TAC AGC-3 0 . For positive controls and normalization b-actin<br />

and Na + /K + -ATPase primer pairs were: Actin/F, 5 0 -GGT CAT CAC TAT CGG<br />

CAA TGA GC-3 0 ; Actin/R, 5 0 -GGA CAG TGA GGC CAG GAT AGA GC-3 0 ;<br />

ATPase/F, 5 0 -AGT GAG CTG AAA CCC ACG TAC C-3 0 and ATPase/R, 5 0 -<br />

CCC CTC TTT GTA GCC GTA GGA TT-3 0 . The resulting PCR products were<br />

cloned into pGMT vector (Promega, Mannheim, Germany) and sequenced.<br />

3. Results<br />

In a previous study <strong>of</strong> olfactory sensory neurons, we<br />

observed that these neurons were incapable <strong>of</strong> maintaining<br />

elevated [Cl ] i after cell isolation (Kaneko et al., 2004). In the<br />

present study we, therefore, measured [Cl ] i without isolating<br />

the somatosensory neurons from their ganglia. Moreover, the<br />

protein expression pattern in isolated somatosensory neurons<br />

changes during dedifferentiation processes in culture (Scott and<br />

Edwards, 1980), and this may alter the control <strong>of</strong> [Cl ] i .To


Author's personal copy<br />

482<br />

D. Gilbert et al. / Int. J. Devl Neuroscience 25 (2007) 479–489<br />

obtain information about [Cl ] i values in situ, we performed<br />

<strong>chloride</strong> imaging in dorsal root ganglia with intact dura mater<br />

and intact axonal connections with the spinal cord. Isolated<br />

neurons were only used for calibrating the 2P-FLIM signal, as<br />

the manipulation <strong>of</strong> [Cl ] i by ionophores was more efficient<br />

when the neurons were dissociated.<br />

3.1. Somatosensory neurons undergo a postnatal <strong>chloride</strong><br />

transition<br />

The neurons in MQAE-loaded dorsal root ganglia displayed<br />

different fluorescence intensities when illuminated with the<br />

excitation light (Fig. 1a, black-and-white image). The fluorescence<br />

intensity, however, does not report [Cl ] i because it is<br />

co-determined by the unknown dye concentration in each cell.<br />

Fluorescence originating from the nuclei was weak. This may<br />

be caused by poor dye loading <strong>of</strong> the nuclei or by different<br />

quenching conditions in the nucleoplasm. The nuclear signals<br />

where excluded from further analyses. The fluorescence<br />

lifetime, t, which is proportional to [Cl ] i and can be used<br />

as intracellular Cl reporter (Koncz and Daugirdas, 1994;<br />

Kaneko et al., 2002), was colour-coded and is displayed as 2P-<br />

FLIM image for the same cells (Fig. 1a, colour image). In the<br />

2P-FLIM images, warmer colours indicate higher levels <strong>of</strong><br />

[Cl ] i . To establish the quantitative relation between 2P-FLIM<br />

signals and [Cl ] i in MQAE-loaded dorsal root ganglion<br />

neurons, we set [Cl ] i in isolated neurons to 1–5 levels (15, 30,<br />

45, 60, 75 mM) using a double ionophore technique. Isolated<br />

neurons in primary culture were used because the manipulation<br />

<strong>of</strong> [Cl ] i was more efficient in single cells than in intact ganglia.<br />

The neurons were kept in a solution containing the required<br />

Cl concentration (Table 1) as well as 40 mM tributyltin (a Cl /<br />

OH exchanger) and 20 mM nigericin (a K + /H + exchanger).<br />

The combination <strong>of</strong> these ionophores has been shown to<br />

dissipate Cl gradients across the plasma membrane (Chao<br />

et al., 1989). Fluorescence lifetimes decreased from 3.7 to<br />

3.1 ns when [Cl ] i was raised from 15 to 75 mM within the<br />

cytosol (15 mM: 3.73 0.17 ns, n = 16; 30 mM: 3.66 <br />

0.09 ns, n = 4; 45 mM: 3.45 0.14 ns, n =5; 60mM:<br />

3.43 0.07 ns, n = 13; 75 mM: 3.14 0.13 ns, n = 10). A<br />

Stern–Volmer plot <strong>of</strong> the calibration data (Fig. 1b) yielded a<br />

quenching constant <strong>of</strong> 3.05 M 1 , which was used in all further<br />

experiments to calculate [Cl ] i from the measured lifetimes.<br />

We have previously tested the reliability <strong>of</strong> this calibration<br />

procedure in primary afferent neurons <strong>of</strong> the olfactory system.<br />

In a comparison <strong>of</strong> 2P-FLIM analysis (Kaneko et al., 2004) and<br />

<strong>chloride</strong> measurements by energy-dispersive X-ray microanalysis<br />

(Reuter et al., 1998), the measured [Cl ] i values differed<br />

Fig. 1. Determination <strong>of</strong> [Cl ] i in somatosensory neurons by 2P-FLIM. (a) Comparison <strong>of</strong> fluorescence intensity (left) and lifetime (right) images from the same<br />

dorsal root ganglion. The intensity image shows strong signals from the cytosol and weak signals from the nuclei. In the 2P-FLIM image, the fluorescence lifetime, t,<br />

is colour-coded with warmer colours representing higher [Cl ] i levels. (b) Calibration <strong>of</strong> 2P-FLIM signals in isolated dorsal root ganglion neurons with [Cl ] i set to<br />

the indicated values using the two-ionophore method. The solid line represents a least-squares fit to the data using the Stern–Volmer equation t 0 /t =1+K SV [Cl ] i<br />

with K SV = 3.05 M 1 . The colour scale illustrates the false-colour representation <strong>of</strong> [Cl ] i in the following 2P-FLIM images. (c) 2P-FLIM images illustrating [Cl ] i<br />

levels in somatosensory neurons from newborn (P1–P4) and adult (third week) mice. The calibration established in (b) was used for false-colour representation <strong>of</strong><br />

[Cl ] i . Newborn neurons show almost uniformly high [Cl ] i (70 mM). During <strong>maturation</strong>, most somatosensory neurons decrease their [Cl ] i to some extent,<br />

resulting in a heterogeneous mosaic <strong>of</strong> [Cl ] i levels in the ganglia <strong>of</strong> mature animals.


Author's personal copy<br />

D. Gilbert et al. / Int. J. Devl Neuroscience 25 (2007) 479–489 483<br />

by less than 20%. For this reason, we give all absolute Cl<br />

concentrations in the present study with a confidence range<br />

<strong>of</strong> 20%.<br />

We determined [Cl ] i in somatosensory neurons in intact<br />

ganglia by focusing the excitation light <strong>of</strong> the 2P-FLIM<br />

instrument through the dura mater into cells 20–100 mm deep<br />

inside the ganglion. Deeper cell layers were not accessible for<br />

analysis because the dye did not sufficiently load the cells at<br />

distances > 100 mm from the dura mater. To determine [Cl ] i<br />

in neurons <strong>of</strong> newborn mice, preparations <strong>of</strong> 1–4 days old<br />

animals (P1–P4) were examined by 2P-FLIM. Since the<br />

<strong>chloride</strong> transition develops between P6 and P14 (Rivera et al.,<br />

2005), we expected to find in the P1–P4 animals the high levels<br />

<strong>of</strong> [Cl ] i characteristic for embryonal and early postnatal<br />

neurons. Indeed, the somatosensory neurons exhibited uniformly<br />

high levels <strong>of</strong> [Cl ] i (Fig. 1c, P1–P4). Single cell-based<br />

quantitative analysis <strong>of</strong> 975 neurons (pooled from P1 to P4)<br />

yielded a distribution <strong>of</strong> [Cl ] i between 35 and 120 mM with a<br />

mean [Cl ] i <strong>of</strong> 77.2 mM (S.D. = 13.6 mM) (Fig. 2).<br />

2P-FLIM measurements in the third postnatal week revealed a<br />

decrease <strong>of</strong> [Cl ] i in most cells, and a heterogeneous distribution<br />

<strong>of</strong> [Cl ] i levels among different neurons, ranging from 10 to<br />

>80 mM [Cl ] i (Fig. 1c). The general decline <strong>of</strong> [Cl ] i indicates<br />

a pr<strong>of</strong>ound change in <strong>chloride</strong> <strong>homeostasis</strong> with reduced Cl<br />

uptake and/or enhanced Cl extrusion. In adult animals (12–15<br />

weeks), the heterogeneous distribution was maintained (Fig. 1c).<br />

A histogram <strong>of</strong> [Cl ] i values obtained from 818 neurons in eight<br />

ganglia from six adult animals is shown in Fig. 2. The mean value<br />

is significantly lower than in P1–P4 animals (Student’s t-test:<br />

p 0.0005), and the distribution is broad (mean: 61.8 mM,<br />

S.D. = 19.7 mM). Both postanatal and adult data could be<br />

Fig. 2. Distribution <strong>of</strong> [Cl ] i values in dorsal root ganglia <strong>of</strong> newborn and adult<br />

mice. The histograms show the distribution <strong>of</strong> measured [Cl ] i levels in<br />

newborn (closed circles) and adult (open circles) animals. During the course<br />

<strong>of</strong> <strong>maturation</strong>, the distribution <strong>of</strong> [Cl ] i shifts towards lower levels. The lines<br />

compare global fits to the data with a single Gaussian distribution (dashed lines)<br />

and with three Gaussian distributions (solid lines). The three populations<br />

contributing to the latter fit are characterized by mean [Cl ] i levels <strong>of</strong><br />

37.8 mM, 58.7 mM, and 80.5 mM, and their relative contributions change<br />

during <strong>maturation</strong> from 0 to 18.9%, from 21 to 55.2%, and from 79 to<br />

25.9%, respectively.<br />

tentatively described by a sum <strong>of</strong> three Gaussian distributions<br />

with mean values <strong>of</strong> 37.8 mM, 58.7 mM and 80.5 mM [Cl ] i ,<br />

respectively (Fig. 2). During the developmental transition, the<br />

percentage <strong>of</strong> ‘‘high-<strong>chloride</strong> cells’’ decreased from 79 to 26%,<br />

that <strong>of</strong> ‘‘medium-<strong>chloride</strong> cells’’ increased from 21 to 55%, and<br />

that <strong>of</strong> ‘‘low-<strong>chloride</strong> cells’’ from 0 to 19%. Our data demonstrate<br />

that the somatosensory neurons in an individual ganglion <strong>of</strong> an<br />

adult animal exhibit a differential pattern <strong>of</strong> <strong>chloride</strong> <strong>homeostasis</strong>.<br />

Most neurons undergo a transition <strong>of</strong> the <strong>chloride</strong><br />

<strong>homeostasis</strong> during <strong>maturation</strong>, while roughly a third <strong>of</strong> the cells<br />

maintain the high [Cl ] i levels <strong>of</strong> the early postnatal days.<br />

3.2. Lack <strong>of</strong> correlation between [Cl ] i and cell size<br />

The size <strong>of</strong> somatosensory neurons can be indicative <strong>of</strong> their<br />

sensory modalities. Small and medium-sized somata (B-cells;<br />

mean volume: 10,700 mm 3 ; Tandrup, 2004) are <strong>of</strong>ten polymodal<br />

nociceptors or heat-sensitive cells with unmyelinated C-<br />

fibres. Large-diameter neurons (A-cells; mean volume:<br />

57,200 mm 3 ; Tandrup, 2004) mediate proprioception, touch<br />

or other low-threshold sensations. While detailed information<br />

about protein expression patterns, conduction velocities, and<br />

excitation properties is required to establish the specific<br />

modality <strong>of</strong> a somatosensory neuron (Guo et al., 1999; Lawson,<br />

2002; Fang et al., 2006), the cell size can serve as a first<br />

indication for the sensory system involved. To find out whether<br />

the different [Cl ] i levels in adult neurons correspond to any<br />

specific cell size, we measured the diameters <strong>of</strong> all neurons<br />

analyzed for Fig. 2, and related them to the measured [Cl ] i<br />

values. Since the 2P-FLIM images represent optical slices <strong>of</strong><br />

about 1 mm thickness, an individual image may show either the<br />

maximal (equatorial) diameter <strong>of</strong> a cell or, alternatively, a<br />

smaller (non-equatorial) diameter. For our analysis, we<br />

determined the maximal diameter for each individual cell.<br />

[Cl ] i was determined in the cytosolic region <strong>of</strong> each neuron,<br />

excluding the nucleus (Fig. 3a, arrows). The plot <strong>of</strong> [Cl ] i<br />

versus cell diameter does not reveal any correlation between the<br />

two parameters in adult neurons (Fig. 3b, black dots; linear<br />

regression analysis: r 2 = 0.019, n = 818). Both small-diameter<br />

cells (15–25 mm) and larger neurons cover the entire range <strong>of</strong><br />

measured [Cl ] i levels. The mean values (Fig. 3c) illustrate an<br />

increase <strong>of</strong> the mean cell diameter from 17.5 2.8 to<br />

25.8 6.3 mm and a decrease <strong>of</strong> mean [Cl ] i from<br />

77.2 13.6 to 61.8 19.7 mM during <strong>maturation</strong>. However,<br />

the <strong>chloride</strong> transition appears not to be restricted to a cell<br />

population with a distinct size. This finding suggests that a<br />

change in the regime <strong>of</strong> <strong>chloride</strong> handling occurs during<br />

postnatal <strong>maturation</strong> in somatosensory neurons <strong>of</strong> various<br />

sensory modalities.<br />

3.3. Expression <strong>of</strong> NKCC1 and KCC2 proteins<br />

A possible explanation for the mosaic pattern <strong>of</strong> [Cl ] i levels<br />

in adult dorsal root ganglia could be that some cells express<br />

predominantly NKCC1 while others express mainly KCC2. To<br />

test this hypothesis, we double-stained cryosections <strong>of</strong> dorsal<br />

root ganglia with antibodies raised against the two proteins. To


Author's personal copy<br />

484<br />

D. Gilbert et al. / Int. J. Devl Neuroscience 25 (2007) 479–489<br />

Fig. 3. Heterogeneous [Cl ] i levels in mouse somatosensory neurons <strong>of</strong> various sizes. (a) 2P-FLIM images show the nuclear region in many cells with fluorescence<br />

signals distinct from the cytosol. To determine [Cl ] i , the fluorescence lifetime was evaluated in the cytosolic region, illustrated as the area between two circles<br />

(arrows). To determine the cell size, a series <strong>of</strong> images (z-stack) was analyzed, and the maximal diameter <strong>of</strong> each individual cell was measured. (b) Dot plot relating<br />

[Cl ] i to cell diameter for 975 neurons from early postnatal animals (red) and 818 neurons from adult animals (black). No correlation between cell size and [Cl ] i is<br />

detectable. (c) Mean [Cl ] i decreased from 77.2 13.6 to 61.8 19.7 mM, while the mean cell diameter increased from 17.5 2.8 to 25.8 6.3 mm during<br />

<strong>maturation</strong>.<br />

detect NKCC1 expression, we used two different antibodies:<br />

the monoclonal T4 antibody (raised against a C-terminal<br />

epitope <strong>of</strong> NKCC1; Lytle et al., 1995) which was demonstrated<br />

to be appropriate for somatosensory neurons (Alvarez-Leefmans<br />

et al., 2001), and the polyclonal N-16 antibody (raised<br />

against an N-terminal epitope <strong>of</strong> NKCC1, Santa Cruz). Of 534<br />

cells, 450 (84%) were stained with the T4 antibody. Co-staining<br />

with T4 and N-16 antibodies revealed a highly consistent<br />

pattern (Fig. 4a–c). Two hundred and forty-nine small- and<br />

medium-diameter neurons were examined. Eighty-one percent<br />

<strong>of</strong> these (202 cells) were positive for both antibodies, 5% (12<br />

cells) showed only a T4 signal, and 14% (35 cells) were not<br />

stained. Thus, 94% <strong>of</strong> T4-positive neurons were also N-16<br />

positive, which represents strong evidence for NKCC1<br />

expression in most somatosensory neurons. Immunosignals<br />

from the KCC2 antibody were significantly weaker than those<br />

from NKCC1. Seventy-six percent <strong>of</strong> 285 small-to-medium<br />

sized cells were immunopositive for KCC2, and 91% <strong>of</strong> these<br />

were also stained with the T4 antibody (Fig. 4d–f). In summary,<br />

expression <strong>of</strong> KCC2 appears to be distinctly weaker than that <strong>of</strong><br />

NKCC1. While most <strong>of</strong> the somatosensory neurons coexpressed<br />

NKCC1 and KCC2, a subpopulation <strong>of</strong> 20% <strong>of</strong><br />

the small-to-medium size neurons express NKCC1 without<br />

detectable KCC2 expression.<br />

3.4. Invariant transcription <strong>of</strong> electroneutral cation–<br />

<strong>chloride</strong> transporters during <strong>maturation</strong><br />

The change <strong>of</strong> [Cl ] i during postnatal development may be<br />

the consequence <strong>of</strong> altered expression <strong>of</strong> Cl transporters in the<br />

somatosensory neurons. To test whether an altered transcription<br />

program causes the Cl transition, we examined the mRNA


Author's personal copy<br />

D. Gilbert et al. / Int. J. Devl Neuroscience 25 (2007) 479–489 485<br />

Fig. 4. Expression <strong>of</strong> electroneutral cation–<strong>chloride</strong> co-transporters in rat somatosensory neurons. (a–c) Immunohistochemistry with two antibodies directed against<br />

NKCC1. The immunosignals <strong>of</strong> the antibodies T4 (C-terminal) and N-16 (N-terminal) match, with only few neurons missing the N-16 signal (red cells in c). (d–f)<br />

Expression <strong>of</strong> KCC2 is detectable but weak in most somatosensory neurons. No immunostain was observed after preadsorption <strong>of</strong> the polyclonal antibodies (N-16 and<br />

KCC2) with their respective immunization peptides (not shown). Nuclei are stained blue with DAPI.<br />

expression <strong>of</strong> all nine electroneutral cation–<strong>chloride</strong> cotransporters<br />

present in the rat genome (synonym: solute carrier<br />

family 12, SLC12; review, Gamba, 2005) by semi-quantitative<br />

RT-PCR. We made first-strand cDNA from dorsal root ganglia<br />

<strong>of</strong> newborn (P1) and adult animals. The relative concentrations<br />

<strong>of</strong> the cDNAs in the RT-PCR assays were adjusted using the<br />

highly abundant b-actin and the less abundant Na + /K + -ATPase<br />

as internal standards (Fig. 5a). The cDNAs were amplified with<br />

primers specific for CCC6, CCC9, KCC1-4, NKCC1-2 and<br />

TSC. CCC6 and CCC9 are orphan members <strong>of</strong> the SLC12<br />

family with unknown function, while TSC is the thiazidesensitive<br />

Na + –Cl co-transporter (for gene identification see<br />

legend to Fig. 5). Cycling conditions were adjusted to 28 and 32<br />

cycles. Fig. 5b demonstrates that the mRNA <strong>of</strong> five Cl cotransporters<br />

(CCC6, KCC1, KCC3, KCC4, NKCC1) could be<br />

detected by 28 PCR cycles from both P1 and adult animals.<br />

Fig. 5c indicates that 32 PCR cycles reached the amplification<br />

plateau for the five co-transporters listed above. In addition, a<br />

diagnostic signal <strong>of</strong> CCC9, KCC2 and TSC was observed,<br />

while no NKCC2 signal could be detected. Generating PCR<br />

amplification products under these conditions may be<br />

influenced by template abundance or by primer efficiency.<br />

To distinguish between these alternatives, we amplified KCC2<br />

from normalized brain cDNA, as well as NKCC2 and TSC from<br />

normalized kidney cDNA. Strong signals were detected by 32<br />

PCR cycles (Fig. 5a, right). Since no tissue-specific expression<br />

data were available for CCC9, the respective control was<br />

performed on adult dorsal root ganglia cDNA only and showed<br />

a strong signal with 36 cycles (Fig. 5a, right). These data<br />

indicate that each <strong>of</strong> the primer pairs used in this study<br />

exhibited similar efficiency. Therefore, the yield <strong>of</strong> the PCR<br />

amplification products depended only on the level <strong>of</strong> input<br />

template. All primer pairs were designed to span intron/exon<br />

boundaries, to reveal possible genomic DNA contamination.<br />

None <strong>of</strong> the PCR reactions showed amplification <strong>of</strong> genomic<br />

DNA which would have given rise to correspondingly larger<br />

DNA fragments. For unambiguous identification, all PCR<br />

products were eluted from agarose gels, cloned into the vector<br />

pGMT and sequenced.<br />

In summary, we found high expression levels <strong>of</strong> CCC6,<br />

KCC1, KCC3, KCC4 and NKCC1, much lower expression<br />

levels for CCC9, KCC2 and TSC, while no expression for<br />

NKCC2 mRNA was observed. Most importantly, the mRNA<br />

expression patterns <strong>of</strong> P1 and adult animals could not be<br />

distinguished on the grounds <strong>of</strong> the semi-quantitative PCR data.<br />

These results clearly demonstrate that the <strong>maturation</strong> <strong>of</strong> Cl<br />

<strong>homeostasis</strong> cannot be attributed to changes in the transcription<br />

pattern <strong>of</strong> electroneutral cation–<strong>chloride</strong> co-transporters.<br />

4. Discussion<br />

In trying to assess the role <strong>of</strong> Cl accumulation in<br />

somatosensory signal generation, we have examined the<br />

question whether Cl <strong>homeostasis</strong> changes after birth. We


Author's personal copy<br />

486<br />

D. Gilbert et al. / Int. J. Devl Neuroscience 25 (2007) 479–489<br />

Fig. 5. Expression <strong>of</strong> electroneutral cation–<strong>chloride</strong> co-transporters in dorsal root ganglia (DRG) <strong>of</strong> newborn (P1) and adult rats. Semi-quantitative RT-PCR assays<br />

were used to determine expression levels <strong>of</strong> mRNA transcripts encoding cation–<strong>chloride</strong> co-transporter 9 (CCC9; synonym: solute carrier family 12 member 8,<br />

SLC12A8; GenBank accession NM_153625), cation–<strong>chloride</strong> co-transporter 6 (CCC6; synonyms: cation–<strong>chloride</strong> co-transporter-interacting protein, CIP;<br />

SLC12A9; NM_134405), K–Cl co-transporter 1 (KCC1, SLC12A4; NM_019229), K–Cl co-transporter 2 (KCC2; SLC12A5; NM_134363), K–Cl co-transporter<br />

3 (KCC3; SLC12A6; XM_001066756), K–Cl co-transporter 4 (KCC4; SLC12A7; XM_001060536), Na–K–Cl co-transporter 1 (NKCC1, SLC12A2; NM_031798),<br />

Na–K–Cl co-transporter 2 (NKCC2, SLC12A1; NM_019134), thiazide-sensitive Na–Cl co-transporter (TSC; SLC12A3; NM_019345). The primers were specific for<br />

the individual subtypes <strong>of</strong> co-transporters and were designed to span several exons to control for the possible presence <strong>of</strong> genomic DNA in the RNA preparation.<br />

Cycles <strong>of</strong> amplification are indicated for each reaction. The figures are negative images <strong>of</strong> ethidium bromide-stained agarose gels. DNA marker (M): 100 bp ladder.<br />

(a) The cDNA concentrations were normalized by amplification (16, 20, 24, 28 and 32 cycles) with b-actin and ATPase primers, and 0.5 ml <strong>of</strong> cDNA product was used<br />

per amplification. For positive controls, KCC2, NKCC2 and TSC were amplified from kidney or brain cDNA, and CCC9 was amplified with 36 cycles from dorsal<br />

root ganglia cDNA <strong>of</strong> adult animals. (b) Strong signals were observed for CCC6, KCC1, KCC3, KCC4 and NKCC1 in either P1 or adult animals using 26 cycles <strong>of</strong><br />

amplification. No expression was detected under these conditions for CCC9, KCC2, NKCC2 and TSC. (c) Amplification begins to plateau within 32 cycles for CCC6,<br />

KCC1, KCC3, KCC4 and NKCC1 in P1 and adult animals. Weak expression was detected for CCC9, KCC2 and TSC in cDNA from P1 and adult animals. No<br />

expression was detected for NKCC2.<br />

measured [Cl ] i in the somata <strong>of</strong> dorsal root ganglion neurons<br />

using a preparation with intact ganglia from newborn and adult<br />

animals. We avoided the use <strong>of</strong> cultured neurons for the<br />

determination <strong>of</strong> [Cl ] i because the expression <strong>of</strong> Cl<br />

transporters and, hence the equilibrium level <strong>of</strong> [Cl ] i , may<br />

change during culture. [Cl ] i in vivo may also be co-determined<br />

by the satellite glial cells which closely wrap each neuronal<br />

soma inside the ganglion (Hanani, 2005), and which are lost<br />

upon cell isolation. We utilized the Cl dependence <strong>of</strong> the<br />

fluorescence lifetime <strong>of</strong> MQAE to directly access [Cl ] i .We<br />

found that [Cl ] i levels were uniformly high in newborns. In the<br />

course <strong>of</strong> the first 3 weeks after birth, somatosensory neurons


Author's personal copy<br />

D. Gilbert et al. / Int. J. Devl Neuroscience 25 (2007) 479–489 487<br />

changed their mode <strong>of</strong> Cl handling, and most neurons lowered<br />

their [Cl ] i level to some extent. In adult dorsal root ganglia,<br />

[Cl ] i varied over a wide range <strong>of</strong> concentrations. The<br />

distribution <strong>of</strong> [Cl ] i levels among individual neurons indicated<br />

an E Cl range <strong>of</strong> 70 to 20 mV. Statistical analysis pointed to<br />

the presence <strong>of</strong> three distinct populations <strong>of</strong> neurons with [Cl ] i<br />

levels near 40, 60 and 80 mM, respectively. However, the<br />

interpretation <strong>of</strong> this finding requires further examinations. Our<br />

results suggest that Cl currents are inhibitory in some<br />

somatosensory neurons and excitatory in others.<br />

What is the molecular basis <strong>of</strong> this transition? Our PCR<br />

study shows that five members <strong>of</strong> the SLC12-family <strong>of</strong><br />

electroneutral Cl co-transporters (CCC6, KCC1, 3 and 4,<br />

NKCC1) show robust transcription at P1 which does not change<br />

during <strong>maturation</strong>. The mRNA levels <strong>of</strong> three other family<br />

members (CCC9, KCC2, and TSC) are lower, but persist during<br />

<strong>maturation</strong>. Our results are in accordance with in-situ<br />

hybridization studies on dorsal root ganglia from adult rats.<br />

Strong KCC1 and NKCC1 expression was detected with<br />

riboprobes, while KCC2 mRNA could hardly be detected<br />

(Kanaka et al., 2001). Moreover, using a combination <strong>of</strong> in situ<br />

hybridization and immunohistochemistry, Price et al. (2006)<br />

found that the NKCC1 mRNA displays 50% co-localization<br />

with markers <strong>of</strong> unmyelinated nociceptors (peripherin, CGRP,<br />

TRPV1) with diameters in the range <strong>of</strong> 15–35 mm. Only 10–<br />

20% co-localization was observed with a marker for the<br />

myelinated, low-threshold sensory neurons (N52 protein) <strong>of</strong><br />

somewhat larger size (25–55 mm). Interestingly, NKCC1<br />

mRNA was also detected in the satellite glial cells (Price<br />

et al., 2006), the second major cell population in the dorsal root<br />

ganglia. This observation points to a caveat in the interpretation<br />

<strong>of</strong> our PCR data, as they do not reveal from which cell type the<br />

SLC12 templates originated. Nevertheless, our observation that<br />

mRNA levels do not change during the course <strong>of</strong> the Cl<br />

transition argues against the notion that transcriptional<br />

regulation underlies the change in Cl <strong>homeostasis</strong> in<br />

somatosensory neurons.<br />

Our immunohistochemical results indicate that most<br />

somatosensory neurons are equipped with NKCC1 and, hence,<br />

are capable <strong>of</strong> accumulating intracellular Cl . This interpretation<br />

rests on the results obtained with two antibodies (one N-<br />

terminal, one C-terminal) and is consistent with similar reports<br />

from cat and rat dorsal root ganglia (Alvarez-Leefmans et al.,<br />

2001), as well as with the expression pattern <strong>of</strong> NKCC1 mRNA<br />

(Price et al., 2006). A recent study with different antibodies<br />

showed NKCC1 signals mainly or exclusively in the satellite<br />

glial cells (Price et al., 2006). To our knowledge, the antibodies<br />

in this study were not characterized as rigorously on dorsal root<br />

ganglia as the T4 and N-16 antibodies used here. The<br />

monoclonal T4 antibody stains somatosensory neurons and<br />

satellite glial cells, and it recognizes a single band <strong>of</strong> the<br />

appropriate size in a Western blot from dorsal root ganglia<br />

(Alvarez-Leefmans et al., 2001). Under our experimental<br />

conditions, staining <strong>of</strong> satellite glial cells is discernible, but<br />

staining <strong>of</strong> neurons is prominent. The polyclonal N-16 antibody<br />

appears to be specific as indicated by the preadsorption control.<br />

Together with the matching mRNA expression pr<strong>of</strong>ile (Price<br />

et al., 2006), the immunohistochemical data thus indicate<br />

that somatosensory neurons in situ express NKCC1. This<br />

result is in good agreement with the findings <strong>of</strong> Sung et al.<br />

(2000) who analyzed the reversal voltage <strong>of</strong> GABA-induced<br />

currents in NKCC1 knockout mice (Delpire et al., 1999). In<br />

NKCC1 / mice, the reversal voltage was shifted to more<br />

negative values, indicating a loss <strong>of</strong> Cl accumulation efficacy.<br />

The NKCC1 / animals also showed altered pain behavior,<br />

pointing to an important role <strong>of</strong> NKCC1 in nociceptor<br />

physiology (Sung et al., 2000; Laird et al., 2004; see also<br />

Granados-Soto et al., 2005).<br />

How is [Cl ] i lowered in most adult somatosensory neurons<br />

– in some cells even into the 10 mM range – without change in<br />

the SLC12 transcription pattern? Several recent reports support<br />

the hypothesis that the activity <strong>of</strong> <strong>chloride</strong>–cation cotransporters<br />

is altered by posttranslational modification. Two<br />

distinct mechanisms could be involved: (1) the NKCC1 and<br />

KCC proteins are regulated by phosphorylation (Gagnon et al.,<br />

2006; Giménez, 2006), and this regulation may determine the<br />

[Cl ] i level in each individual neuron. NKCC1 activity is<br />

increased by phosphorylation by the STE20-related protein<br />

kinases SPAK/OSR1 and WNK (Dowd and Forbush, 2003;<br />

Gamba, 2005; Moriguchi et al., 2005; Kahle et al., 2005; Vitari<br />

et al., 2006), while WNK kinases reduce Cl extrusion by<br />

KCC1-4 (De los Heros et al., 2006). Our data indicate that most<br />

somatosensory neurons are equipped with NKCC1 and the four<br />

KCC is<strong>of</strong>orms. Since all <strong>of</strong> these proteins are regulated by<br />

kinases, it is reasonable to speculate that the different [Cl ] i<br />

levels after <strong>maturation</strong> <strong>of</strong> Cl <strong>homeostasis</strong> result from different<br />

phosphorylation states <strong>of</strong> the Cl transporters. (2) The activity<br />

<strong>of</strong> the Cl exporters may depend on oligomerization. A recent<br />

study <strong>of</strong> the <strong>chloride</strong> transition in the auditory brain stem<br />

showed convincingly that the postnatal development <strong>of</strong> Cl<br />

extrusion correlates with oligomerization <strong>of</strong> KCC2 (Blaesse<br />

et al., 2006). Thus, KCC2 appears to be inactive as homomer in<br />

the neurons <strong>of</strong> newborn animals and to be turned active after<br />

dimerization. A similar <strong>maturation</strong> process may develop in the<br />

somatosensory system and may lead to enhanced Cl extrusion.<br />

We did not investigate contributions <strong>of</strong> Cl transporters that<br />

do not belong to the SLC12 family. Cl /HCO 3 exchangers <strong>of</strong><br />

the SLC26 family (Mount and Romero, 2004; Shcheynikov<br />

et al., 2006) may be contributing to [Cl ] i in vivo, or<br />

transporters from the SLC4 family (Romero, 2005). However,<br />

the Cl /HCO 3 exchanger is conceptionally associated with<br />

Cl uptake rather than Cl extrusion, in particular, because<br />

HCO 3 can be continuously generated in the cytosol by<br />

carbonic anhydrase (Rivera et al., 2005). In our experiments,<br />

we did not supply extracellular HCO 3 . It is, therefore, unlikely<br />

that the decrease <strong>of</strong> [Cl ] i in our experiments is caused by Cl<br />

extrusion through Cl /HCO 3 exchange. However, in the<br />

living animal, this transporter may contribute significantly to<br />

setting [Cl ] i levels in somatosensory neurons. Furthermore,<br />

our PCR analysis showed that thiazide-sensitive Na + /Cl cotransporter<br />

mRNA can be detected in dorsal root ganglia, albeit<br />

at a comparably low level as KCC2. It will be interesting to<br />

establish the cellular distribution <strong>of</strong> this protein within the<br />

ganglia.


Author's personal copy<br />

488<br />

D. Gilbert et al. / Int. J. Devl Neuroscience 25 (2007) 479–489<br />

What could be the physiological significance <strong>of</strong> the postnatal<br />

change in Cl <strong>homeostasis</strong> in somatosensory neurons? Recent<br />

studies on peripheral Cl effects indicate that depolarizing Cl<br />

currents amplify the primary signal in the sensory terminals <strong>of</strong><br />

nociceptors. Transduction in these neurons works without<br />

metabotropic amplification <strong>of</strong> the primary signal. The<br />

transduction channels, which are directly gated by heat,<br />

chemical or mechanical stimuli, are invariably Ca 2+ -permeable<br />

(McNaughton, 2004), and most somatosensory neurons express<br />

Ca 2+ -activated Cl channels (Kenyon and G<strong>of</strong>f, 1998; Lee<br />

et al., 2005; Currie et al., 1995). Stimulus-induced Ca 2+ -influx<br />

thus triggers Cl currents, and those neurons which accumulate<br />

Cl in their sensory endings are depolarized by Cl efflux<br />

(Granados-Soto et al., 2005). Accordingly, pain behavior can be<br />

induced in animal models by excitatory Cl currents in the skin<br />

(Ault and Hildebrand, 1994), and can be attenuated by<br />

experimental inhibition <strong>of</strong> Cl accumulation into the sensory<br />

endings (Willis et al., 2004; Laird et al., 2004). Based on these<br />

observations, Price et al. (2005) proposed that increased Cl<br />

accumulation into sensory endings <strong>of</strong> nociceptors promotes<br />

hyperalgesia. Our data suggest that the Cl dependence <strong>of</strong> the<br />

nociceptive response reflects the actual state <strong>of</strong> Cl accumulation<br />

in each individual nociceptive neuron. In general, the<br />

sensitivity <strong>of</strong> somatosensory neurons may be co-determined by<br />

the efficiency <strong>of</strong> Cl accumulation, which, in turn, is regulated<br />

by posttranslational modification <strong>of</strong> its Cl transporters.<br />

Interestingly, a recent study <strong>of</strong> [Cl ] i levels in axotomized<br />

mouse somatosensory neurons revealed that NKCC1 phosphorylation<br />

and a consequent rise <strong>of</strong> [Cl ] i from 31 to 68 mM is<br />

associated with peripheral nerve regeneration (Pieraut et al.,<br />

2007). Thus, a dynamic regulation <strong>of</strong> [Cl ] i levels may be<br />

involved in different cellular processes in somatosensory<br />

neurons, including signal transduction and growth.<br />

Taken together, this study demonstrates that Cl <strong>homeostasis</strong><br />

in somatosensory neurons develops during postnatal<br />

<strong>maturation</strong> into a state where [Cl ] i can be regulated<br />

individually in each neuron. Regulation does not affect the<br />

transcriptional level <strong>of</strong> electroneutral cation–<strong>chloride</strong> cotransporters,<br />

but may exert its control via posttranslational<br />

modification. The efficiency <strong>of</strong> Cl accumulation sets the level<br />

<strong>of</strong> [Cl ] i and may, hence, control the sensitivity <strong>of</strong> the adult<br />

sensory neuron.<br />

Acknowledgements<br />

We thank Dr. Joe Lynch (University <strong>of</strong> Queensland) for<br />

helpful comments on the manuscript. This work was supported<br />

by the Deutsche Forschungsgemeinschaft (Fr 937/6).<br />

References<br />

Alvarez-Leefmans, F.J., Gamino, S.M., Giraldez, F., Nogueron, I., 1988.<br />

Intracellular <strong>chloride</strong> regulation in amphibian dorsal root ganglion neurons<br />

studied with ion-sensitive microelectrodes. J. Physiol. 406, 225–246.<br />

Alvarez-Leefmans, F.J., Leon-Olea, M., Mendoza-Sotelo, J., Alvarez, F.J.,<br />

Anton, B., Garduno, R., 2001. Immunolocalization <strong>of</strong> the Na + -K + -2Cl<br />

cotransporter in peripheral nervous tissue <strong>of</strong> vertebrates. Neuroscience 104,<br />

569–582.<br />

Ault, B., Hildebrand, L.M., 1994. GABA A receptor-mediated excitation <strong>of</strong><br />

nociceptive afferents in the rat isolated spinal cord-tail preparation. Neuropharmacology<br />

33, 109–114.<br />

Ben-Ari, Y., 2002. Excitatory actions <strong>of</strong> GABA during development: the nature<br />

<strong>of</strong> the nurture. Nat. Rev. Neurosci. 3, 728–739.<br />

Bevensee, M.O., Apkon, M., Boron, W.F., 1997. Intracellular pH regulation in<br />

cultured astrocytes from rat hippocampus. J. Gen. Physiol. 110, 467–483.<br />

Blaesse, P., Guillemin, I., Schindler, J., Schweizer, M., Delpire, E., Khirough,<br />

L., Friauf, E., Nothwang, H.G., 2006. Oligorimerization <strong>of</strong> KCC2 correlates<br />

with development <strong>of</strong> inhibitory neurotransmission. J. Neurosci. 26, 10407–<br />

10419.<br />

Chao, A.C., Dix, J.A., Sellers, M.C., Verkman, A.S., 1989. Fluorescence<br />

measurement <strong>of</strong> <strong>chloride</strong> transport in monolayer cultured cells. Biophys.<br />

J. 56, 1071–1081.<br />

Chen, H., Luo, J., Kintner, D.B., Shull, G.E., Sun, D., 2005. Na + -dependent<br />

<strong>chloride</strong> transporter (NKCC1)-null mice exhibit less gray and white<br />

matter damage after focal cerebral ischemia. J. Cer. Blood Flow Metabol.<br />

25, 54–66.<br />

Chomczynski, P., Sacchi, N., 1987. Single-step method <strong>of</strong> RNA isolation by<br />

acid guanidinium thiocyanate-phenol-chlor<strong>of</strong>orm extraction. Anal. Biochem.<br />

162, 156–159.<br />

Currie, K.P.M., Wooton, J.F., Scott, R.H., 1995. Activation <strong>of</strong> Ca 2+ -dependent<br />

Cl currents in cultured sensory neurons by flash photolysis <strong>of</strong> DMnitrophen.<br />

J. Physiol. Lond. 482, 291–307.<br />

De Koninck, Y., 2007. Altered <strong>chloride</strong> <strong>homeostasis</strong> in neurological disorders: a<br />

new target. Curr. Opin. Pharmacol. 7, 93–99.<br />

De los Heros, P., Kahle, K.T., Rinehart, J., Bobadilla, N.A., Vázquez, N., San<br />

Cristobal, P., Mount, D.B., Lifton, R.P., Hebert, S.C., Gamba, G., 2006.<br />

WNK3 bypasses the tonicity requirement for K-Cl cotransport activation<br />

via a phosphatase-dependent pathway. Proc. Natl. Acad. Sci. USA 103,<br />

1976–1981.<br />

Delpire, E., Lu, J., England, R., Dull, C., Thorne, T., 1999. Deafness and<br />

imbalance associated with inactivation <strong>of</strong> the secretory Na-K-2Cl cotransporter.<br />

Nat. Gen. 22, 192–195.<br />

Dowd, B.F.X., Forbush, B., 2003. PASK (proline–alanine-rich STE20-related<br />

kinase), a regulatory kinase <strong>of</strong> the Na-K-Cl cotransporter (NKCC1). J. Biol.<br />

Chem. 278, 27347–27353.<br />

Eberhardson, M., Patterson, S., Grapengiesser, E., 2000. Micr<strong>of</strong>luorimetric<br />

analysis <strong>of</strong> Cl permeability and its relation to oscillatory Ca 2+ signalling in<br />

glucose-stimulated pancreatic b-cells. Cell Signal 12, 781–786.<br />

Fang, X., Djouhri, L., McMullan, S., Berry, C., Waxman, S.G., Okuse, K.,<br />

Lawson, S.N., 2006. Intense isolectin-B4 binding in rat dorsal root ganglion<br />

neurons distinguishes C-fiber nociceptors with broad action potentials and<br />

high Nav1.9 expression. J. Neurosci. 26, 7281–7292.<br />

Frings, S., Reuter, D., Kleene, S.J., 2000. Neuronal Ca 2+ -activated Cl channels<br />

– homing in on an elusive channel species. Prog. Neurobiol. 60, 247–289.<br />

Gagnon, K.B.E., England, R., Delpire, E., 2006. Characterization <strong>of</strong> SPAK and<br />

OSR1, regulatory kinases <strong>of</strong> the Na-K-2Cl cotransporter. Mol. Cell Biol. 26,<br />

689–698.<br />

Gallagher, J.P., Hihashi, H., Nishi, S., 1978. Characterization and ionic basis <strong>of</strong><br />

GABA-induced depolarizations recorded in vivo from rat primary afferent<br />

neurons. J. Physiol. 275, 263–282.<br />

Gamba, G., 2005. Molecular physiology and pathophysiology <strong>of</strong> electroneutral<br />

cation-<strong>chloride</strong> cotransporters. Physiol. Rev. 85, 423–493.<br />

Giménez, I., 2006. Molecular mechanisms and regulation <strong>of</strong> furosemidesensitive<br />

Na-K-Cl cotransporters. Curr. Opin. Nephrol. Hypertens. 15,<br />

517–523.<br />

Granados-Soto, V., Arguelles, C.F., Alvarez-Leefmans, F.J., 2005. Peripheral<br />

and central antinociceptive action <strong>of</strong> Na + -K + -2Cl cotransporter blockers<br />

on formalin-induced nociception in rats. Pain 114, 231–238.<br />

Guo, A., Vulchanova, L., Wang, J., Li, X., Elde, R., 1999. Immunocytochemical<br />

localization <strong>of</strong> vanilloid receptor 1 (VR1): relationship to<br />

neuropeptides, the P2X3 purinoceptor and IB4 binding sites. Eur. J.<br />

Neurosci. 11, 946–958.<br />

Hanani, M., 2005. Satellite glial cells in sensory ganglia: from form to function.<br />

Brain Res. Rev. 48, 457–476.<br />

Kahle, K.T., Rinehart, J., de los Heros, P., Louvi, A., Meade, P., Vazquez, N.,<br />

Hebert, S.C., Gamba, G., Gimenez, I., Lifton, R., 2005. WNK3 modulates


Author's personal copy<br />

D. Gilbert et al. / Int. J. Devl Neuroscience 25 (2007) 479–489 489<br />

transport <strong>of</strong> Cl in and out <strong>of</strong> cells: implications for control <strong>of</strong> cell volume<br />

and neuronal excitability. Proc. Natl. Acad. Sci. USA 102, 16783–16788.<br />

Kanaka, C., Ohno, K., Okabe, A., Kuriyama, K., Itoh, T., Fukuda, A., Sato, K.,<br />

2001. The differential expression patterns <strong>of</strong> messenger RNAs encoding K-<br />

Cl cotransporters (KCC1,2) and Na-K-2Cl cotransporter (NKCC1) in the rat<br />

nervous system. Neuroscience 104, 933–946.<br />

Kaneko, H., Putzier, I., Frings, S., Gensch, T., 2002. Determination <strong>of</strong> intracellular<br />

<strong>chloride</strong> concentration in dorsal root ganglion neurons by fluorescence<br />

lifetime imaging. In: Fuller, C.M. (Ed.), Calcium-Activated Chloride<br />

Channels. Academic Press, Boston, pp. 167–189.<br />

Kaneko, H., Putzier, I., Frings, S., Kaupp, U.B., Gensch, T., 2004. Chloride<br />

accumulation in mammalian olfactory sensory neurons. J. Neurosci. 24,<br />

7931–7938.<br />

Kenyon, J.L., 2001. The reversal potential <strong>of</strong> Ca 2+ -activated Cl currents<br />

indicates that chick sensory neurons accumulate intracellular Cl . Neurosci.<br />

Lett. 296, 9–12.<br />

Kenyon, J.L., G<strong>of</strong>f, H.R., 1998. Temperature-dependence <strong>of</strong> Ca 2+ current, Ca 2+ -<br />

activated Cl current and Ca 2+ transients in sensory neurons. Cell Calcium<br />

24, 35–48.<br />

Koncz, C., Daugirdas, J.T., 1994. Use <strong>of</strong> MQAE for measurements <strong>of</strong> intracellular<br />

[Cl ] in cultured aortic smooth muscle cells. Am. J. Physiol. 267,<br />

H2114–H2123.<br />

Laird, J.M.A., García-Nicas, E., Delpire, E.J., Cervero, F., 2004. Presynaptic<br />

inhibition and spinal pain processing in mice: a possible role for the<br />

NKCC1 cation-<strong>chloride</strong> co-transporter in hyperalgesia. Neurosci. Lett.<br />

361, 200–203.<br />

Lakowicz, J.R., 1999. Principles <strong>of</strong> Fluorescence Spectroscopy, second ed.<br />

Kluwer Academic/Plenum, New York, London.<br />

Lau, K.R., Evans, R.L., Case, R.M., 1994. Intracellular Cl concentration in<br />

striated intralobular ducts from rabbit mandibular salivary glands. Plügers<br />

Arch. – Eur. J. Physiol. 427, 24–32.<br />

Lawson, S.N., 2002. Phenotype and function <strong>of</strong> somatic primary afferent<br />

nociceptive neurons with C-, Ad or Aa/b-fibres. Exp. Physiol. 87, 239–244.<br />

Lee, M.G., MacGlashan, D.W., Undem, B.J., 2005. Role <strong>of</strong> <strong>chloride</strong> channels in<br />

bradykinin-induced guinea pig airway vagal C-fibre activation. J. Physiol.<br />

566, 205–212.<br />

Lu, J., Karadsheh, M., Delpire, E., 1999. Developmental regulation <strong>of</strong> the<br />

neuronal-specific is<strong>of</strong>orm <strong>of</strong> K-Cl cotransporter KCC2 in postnatal rat<br />

brains. J. Neurobiol. 39, 558–568.<br />

Lytle, C., Xu, J.-C., Biemesderfer, D., Forbush, B., 1995. Distribution and<br />

diversity <strong>of</strong> Na-K-Cl cotransport proteins: a study with monoclonal antibodies.<br />

Am. J. Physiol. 269, C1496–C1505.<br />

Maglova, L.M., Crowe, W.E., Smith, P.R., Altamirano, A.A., Russel, J.M.,<br />

1998. Na + -K + -Cl cotransport in human fibroblasts is inhibited by cytomegalovirus<br />

infection. Am. J. Physiol. 275, C1330–C1341.<br />

McNaughton, P.A., 2004. Pain transduction: gating and modulation <strong>of</strong> ion<br />

channels. In: Transduction Channels in Sensory Cells, Wiley-VCH, Weinheim,<br />

pp. 251–270.<br />

Moriguchi, T., Urushiyama, S., Hisamoto, N., Iemura, S., Uchida, S., Natsume,<br />

T., Matsumoto, K., Shibuya, H., 2005. WNK1 regulates phosphorylation <strong>of</strong><br />

cation-<strong>chloride</strong>-coupled cotransporters via the STE20-related kinases,<br />

SPAK and OSR1. J. Biol. Chem. 280, 42685–42693.<br />

Mount, D.B., Romero, M.F., 2004. The SLC26 gene family <strong>of</strong> multifunctional<br />

anion exchangers. Pflügers Arch – Eur. J. Physiol. 447, 710–721.<br />

Payne, J.A., Rivera, C., Voipio, J., Kaila, K., 2003. Cation-<strong>chloride</strong> co-transporters<br />

in neuronal communication, development and trauma. Trends<br />

Neurosci. 26, 199–206.<br />

Petruska, J.C., Napaporn, J., Johnson, R.D., Cooper, B.Y., 2002. Chemical<br />

responsiveness and histochemical phenotype <strong>of</strong> electrophysiologically<br />

classified cells <strong>of</strong> the adult rat dorsal root ganglion. Neuroscience 115,<br />

15–30.<br />

Pieraut, S., Laurent-Matha, V., Sar, C., Hubert, Th., Méchaly, I., Hilaire, C.,<br />

Mersel, M., Delpire, E., Valmier, J., Scamps, F., 2007. NKCC1 phosphorylation<br />

stimulates neurite growth <strong>of</strong> injured adult sensory neurons. J.<br />

Neurosci. 27, 6751–6759.<br />

Pond, B.B., Berglund, K., Kuner, T., Feng, G., Augustine, G.J., Schwartz-<br />

Bloom, R.D., 2006. The <strong>chloride</strong> transporter Na + -K + -Cl cotransporter<br />

is<strong>of</strong>orm-1 contributes to intracellular <strong>chloride</strong> increases after in vivo<br />

ischemia. J. Neurosci. 26, 1396–1406.<br />

Price, T.J., Cervero, F., de Koninck, Y., 2005. Role <strong>of</strong> cation-<strong>chloride</strong>-cotransporters<br />

(CCC) in pain and hyperalgesia. Curr. Top. Med. Chem. 5, 547–<br />

555.<br />

Price, T.J., Hargreaves, K.M., Cervero, F., 2006. Protein expression and mRNA<br />

cellular distribution <strong>of</strong> the NKCC1 cotransporter in the dorsal root and<br />

trigeminal ganglia <strong>of</strong> the rat. Brain Res. 1112, 146–158.<br />

Reuter, D., Zierold, K., Schröder, W.H., Frings, S., 1998. A depolarizing<br />

<strong>chloride</strong> current contributes to chemoelectrical transduction in olfactory<br />

sensory neurons in situ. J. Neurosci. 18, 6623–6630.<br />

Rivera, C., Voipio, J., Payne, J.A., Ruusuvuori, E., Lahtinen, H., Lamsa, K.,<br />

Pirvola, U., Saarma, M., Kaila, K., 1999. The K + /Cl co-transporter KCC2<br />

renders GABA hyperpolarizing during neuronal <strong>maturation</strong>. Nature 397,<br />

251–255.<br />

Rivera, C., Voipio, J., Kaila, K., 2005. Two developmental switches in<br />

GABAergic signalling: the K + -Cl cotransporter KCC2 and carbonic<br />

anhydrase CAVII. J. Physiol. 562, 27–36.<br />

Romero, M.F., 2005. Molecular pathophysioloy <strong>of</strong> SLC4 bicarbonate transporters.<br />

Curr. Opin. Nephrol. Hypertens. 14, 495–501.<br />

Rudomin, P., Schmidt, R.F., 1999. Presynaptic inhibition in the vertebrate spinal<br />

cord revisited. Exp. Brain Res. 129, 1–37.<br />

Scott, B.S., Edwards, B.A., 1980. Electric membrane properties <strong>of</strong> adult mouse<br />

DRG neurons and the effect <strong>of</strong> culture duration. J. Neurobiol. 11, 291–301.<br />

Shcheynikov, N., Wang, Y., Meeyoung, P., Ko, S.B.H., Dorwart, M., Naruse, S.,<br />

Thomas, P.J., Muallem, S., 2006. Coupling modes and stoichiometry <strong>of</strong> Cl /<br />

HCO 3 exchange by slc26a3 and slc26a6. J. Gen. Physiol. 127, 511–524.<br />

Stein, V., Hermans-Borgmeyer, I., Jentsch, T.H., 2004. Expression <strong>of</strong> the KCl<br />

cotransporter KCC2 parallels neuronal <strong>maturation</strong> and the emergence <strong>of</strong><br />

low intracellular <strong>chloride</strong>. J. Comp. Neurol. 468, 57–64.<br />

Sung, K.W., Kirby, M., McDonald, M.P., Lovinger, D.M., Delpire, E., 2000.<br />

Abnormal GABA A receptor-mediated currents in dorsal root ganglion<br />

neurons isolated from Na-K-2Cl cotransporter null mice. J. Neurosci. 20,<br />

7531–7538.<br />

Tandrup, T., 2004. Unbiased estimates <strong>of</strong> number and size <strong>of</strong> rat dorsal root<br />

ganglion cells in studies <strong>of</strong> structure and cell survival. J. Neurocytol. 33,<br />

173–192.<br />

Verkman, A.S., 1990. Development and biological applications <strong>of</strong> <strong>chloride</strong>sensitive<br />

fluorescent indicators. Am. J. Physiol. 259, C375–C388.<br />

Vitari, A.C., Thastrup, J., Rafiqi, F.H., Deak, M., Morrice, N.A., Karlsson,<br />

H.K.R., Alessi, D.R., 2006. Functional interactions <strong>of</strong> the SPAK/OSR1<br />

kinases with their upstream activator WNK1 and downstream substrate<br />

NKCC1. Biochem. J. 397, 223–231.<br />

Willis, E.F., Clough, G.F., Church, M.K., 2004. Investigation into the mechanisms<br />

by which nedocromil, furosemide and bumetanide inhibit the histamininduced<br />

itch and flare response in human skin in vivo. Clin. Exp. Allergy 34,<br />

450–455.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!