Skip to main content
Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Ann Neurosci. 2011 Apr; 18(2): 70–75.
PMCID: PMC4117038
PMID: 25205925

Emerging role of WNK1 in pathologic central nervous system signaling

Abstract

WNK1 (with no lysine (K)) is a widely expressed serine/threonine protein kinase. The role of this kinase was first described in the kidney where it dynamically controls ion channels that regulate changes in cell volume. WNK1, through intermediates oxidative stress-responsive kinase-1 (OSR1) and STE20/SPS1-related proline/alanine-rich kinase (SPAK), phosphorylates the inwardly directed Na+-K+-Cl---cotransporter 1 (NKCC1) and the outwardly directed K+-Cl--cotransporter 2 (KCC2), activating and deactivating these channels, respectively. WNK1, NKCC1 and KCC2 are also expressed in the central nervous system (CNS). Growing evidence implicates WNK1 playing a critical role in pathologic nervous system signaling where changes in intracellular ion concentration in response to γ-aminobutyric-acid (GABA) can activate otherwise silent pathways. This review will focus on current research about WNK1, its downstream effectors and role in GABA signaling. Future perspectives include investigating WNK1 expression in the CNS after spinal cord injury (SCI), where altered neuronal signaling could underlie pathological states such as neuropathic pain (NP).

Keywords: NKCC1, KCC2, GABA, Neuropathic Pain, Spinal Cord Injury

Introduction

The NKCC1 and KCC2 Channels

The human form of the Na+-K+-Cl cotransporter 1 (NKCC1) channel is located on chromosome 5q.23.2 and is expressed by the SLC12a2 gene. This cotransporter is a 1212 amino acid protein with a molecular weight of 132 kDa and 12 transmembrane domains. The NKCC1 channel transports 1 Na+:1 K+:2 Cl- into the cell.1 This channel is highly expressed on the apical surface of mammalian neurons in the mature central nervous system2-4 and dorsal root ganglion (DRG) sensory neurons in the peripheral nervous system.3,5

The K+-Cl--cotransporter 2 (KCC2) protein, also expressed by the SLC12 gene,6 contains 1116 amino acids, 12 transmembrane domains and has a molecular mass of 123.6 kDa. It transports potassium and chloride out of the cell in 1:1 stoichiometry.7 The channel is neuronal specific8,9 and is found primarily in dendritic spines of inhibitory synapses in the dorsal horn of the spinal cord.9,10

The SLC12 channels may play a role in epilepsy and pathological excitability. Bumetanide (BU), a NKCC1 blocker, suppresses seizures and attenuates electrographic activity in neonatal rats, in vivo.11 Similarly, mice lacking KCC2 channels frequently seize and die shortly after birth.9,12 Three hours of epileptic-like neuronal stress decreases KCC2 mRNA expression in rat hippocampal slices.13

NKCC1 and KCC2 are co-expressed in specific neurons.14,15 After contusion spinal cord injury (cSCI), NKCC1 and KCC2 channel expression are increased and decreased, respectively.16 Phosphorylation activates NKCC1 but inhibits KCC2, whereas dephosphorylation activates KCC2 and deactivates NKCC1.4,17-20

GABA, Chloride and the CNS

Central nervous system (CNS) excitability and behavior is dynamically regulated by variations in intracellular ion concentration.21-23 Changes in [Cl-]i govern the response to the neurotransmitter γ-aminobutyric-acid (GABA). In GABAergic stimulated neurons, Cl-in will occur if [Cl-]i is below ECl-, increasing the probability of hyperpolarization. Conversely if [Cl-]i is above ECl-, GABA stimulation will result in Cl-out, driving the Vm towards ECl-, and potential depolarization.6,22,24 Rat hippocampal slices with downregulated KCC2 channels show reduced Cl- extrusion.25 In KCC2 knockout (K0) mice motorneurons, stimulation with GABA results in depolarization whereas wild-type (WT) neurons hyperpolarize under identical stimulation.9 Spatial-temporal changes in [Cl-]i modify GABA-ergic responses in retinal bipolar cells.14 Early postnatal GABA induced depolarization25 may be due to increased accumulation of [Cl-]i through increased NKCC1 channel expression26,27 and activity.28 As the CNS matures, NKCC1 channel expression is decreased26,29 and KCC2 channel expression is increased;29,30 this could contribute to the changes in [Cl-]i27-29 and subsequent switch of GABA from an excitatory to inhibitory neurotransmitter in development.13,24,27,29-31

WNK Family

WNK1 is a serine/threonine protein kinase that is activated by phosphorylation and was first described in 2000 by Xu et al. as a 2126 amino acid long protein with a molecular weight of about 230 kDa. WNK1 is named so (with no lysine (K)) because it lacks a catalytic lysine found in subdomain II of most of the other protein kinases.32 The WNK1 gene is under the control of at least 3 different promoter regions. This allows for tissue specific distribution. Alterative splicing and polyadenylation sites can achieve further differentiation.33 The WNK1 kinase is found among other places, in cell bodies of DRG neurons.34

Most of the research on the WNK kinases has been done in the kidney and their role in governing blood pressure. Pseudohypoaldosteronism type II (PHAII) is an autosomal dominant disorder where patients present with hypertension and hyperkalemia. Rats with mutations in WNK1 intron 1, mimic PHAII and show a five-fold increase in WNK1 expression. Thus WNK1 appears to play a role in this disease by either increasing the reabsorption of potassium and other ions, or by inhibiting their secretion or excretion.35

The WNK family is upstream activators of the NKCC1 and KCC2 channels. Hypertonic stress increases WNK1 activity.36 WNK1 has been shown to phosphorylate and activate oxidative stress-responsive kinase-1 (OSR1) and STE20/SPS1-related proline/alanine-rich kinase (SPAK, or PASK or STK39).6,37-41 SPAK and OSR1 share amino acid sequence homology in their N-terminal catalytic domain (96%) and C-terminal regulatory domain (67%).41 SPAK and OSR1 phosphorylate three residues on the NKCC1 channel.7,17,38,40-42

Hyperosmotic stress increases NKCC1 phosphorylation17,38 and K+ uptake by this channel.43 WNK1 induced phosphorylation of OSR1 activates this kinase to phosphorylate its NKCC1 substrates in HeLa cells in vitro. HeLa cells injected with WNK1 siRNA exhibit reduced NKCC1 activity.37 MDCKII cells overexpressing WNK1 show increased chloride permeability, in vivo.44 WNK4 phosphorylates SPAK at sites homologous to those phosphorylated by WNK1.39,43 In Xenopus laevis oocytes, coexpression of both WNK4 and SPAK increases NKCC1 channel activation and desensitizes the channel to osmotic conditions. Coexpression of WNK4 and SPAK results in downregulation of KCC2, regardless of osmotic environmental conditions.43 Expression of WNK3 phosphorylates NKCC1 regardless of the osmotic state of the environment. WNK3 increases Cl- influx via NKCC1 and decreases Cl- efflux via the KCC2 channel.17

Thus WNK1, WNK3 and WNK4 behave like volume sensitive kinases that control SLC12 family members. However WNK3 can regulate the NKCC1 and KCC2 transporters alone, where as WNK1/4 require OSR1 and SPAK coexpression; suggesting a separate mechanism for the different kinases. Perhaps WNK3 works by inhibiting phosphatases and thus increasing the phosphorylation state of SCL12 family channels; a different mechanism could exist for WNK1/4: they phosphorylate OSR1 and SPAK which go onto phosphorylate NKCC1 and KCC2.45

The WNK family and its biological cascade play an important role in the nervous system. WNK1 knockdown C17.2 cells show altered morphology, slower motility and reduced invasive ability; suggesting WNK1’s role in proliferation, migration and differentiation in neural development.46 SPAK and OSR1 are expressed in adult neurons of the spinal cord, DRG and brain.47-49 SPAK or OSR1 knockdown mice show about a 50% reduction of spinal cord NKCC1 channel activity,49 and knockdown of both kinases is additive.4 WNK3 is highly expressed in the nervous system and appears to be important in neuronal development: absent in mice on postnatal day 10, but becoming highly expressed by postnatal day 21. This might suggest a role of WNK3 in switching from normal GABA excitation in prenatal life, to GABA inhibition in adulthood.8 Post-mortem analysis of human brain specimens shows schizophrenics have increased WNK3 expression in the dorsolateral prefrontal cortex, an area known to have altered synchrony in diseased patients.50 Additionally, perhaps the WNK family’s ability to dynamically regulate Cl- channels plays a role in the circadian variation of [Cl-]i21,22 and GABA transmission that occurs in the suprachiasmatic nucleus that controls sleep wake cycles.22,51

Hereditary sensory and autonomic neuropathy Type 2 (HSAN2) is a recessive disorder associated with loss of sensitivity.52 A mutated alternatively spliced exon of the WNK1 gene that selectively occurs in nervous tissues called HSN2 is involved in HSAN2.34,52 Specific isoforms of WNK1 have been characterized to be organ specific.53 HSN2 is found primarily in the spinal cord, but is also present in the DRG and sciatic nerve of adult mice. Within the spinal cord, HSN2 is more predominant in the dorsal roots compared to the ventral roots. It is also highly expressed in the laminae II and III, dorsolateral funiculus and lateral funiculus that contain ascending sensory fibers.34 Interestingly, twenty-five human carriers of the defective exon were shown to have lower warm and cool detection thresholds. It was hypothesized that one truncated copy of the WNK1/HSN2 gene results in an increase in membrane excitability lowering detection threshold; however in homozygous HSN2 isoform carriers that have HSAN2, the increased excitability may lead to excitotoxicity leading to decreased sensation.52

Pain Perception

Modulation and/or modification of the nervous system can lead to hyperalgesia (noxious stimuli eliciting a greater than normal pain response) or allodynia (stimuli that normally do not produce pain begin to do so). Recent pain theories propose lose of inhibition (disinhibition) as being crucial for the development of chronic pain.54 There are two types of afferent fibers in the spinal cord: Aβ-fibers that perceive tactile sensations, and Aδ- and C-fibers involved in nociception. A presynaptic link exists between these two fibers55,56 that contains a GABA-ergic interneuron.57 Under normal conditions mechanically stimulated Aβ fibers, acting via the GABA interneuron, will cause primary afferent depolarization (PAD) of the nociceptive terminals; thus shunting pain perception via a presynaptic inhibition mechanism. Following injury, an increased afferent barrage from the Aδ- and C-fibers converges onto the GABA-ergic spinal interneurons that mediate the presynaptic link between mechano and nociceptive receptors. Thus when the Aδ-fibers are now stimulated, the increased excitability of the interneuron produces a much more intense PAD capable of producing spike activity. This results in antidromically conducted dorsal root reflexes (DRR)58 and produce secondary hyperalgesia or allodynia.55,56

Role of NKCC1 and KCC2 in Pain

DRG NKCC1 KO mice show increased thermal pain thresholds. Mutant cells hyperpolarize and WT cells depolarize to identical stimuli, and mutant cells lack the GABAa R-mediated anion outward flux current.5 Blocking NKCC1 channels lowers [Cl-]i accumulation after vagal motorneuron axonotomies.15 Elevations in [Cl-]i after rat sciatic nerve axonotomies is attributable to phosphorylation of the NKCC1 channel.59 Additionally, axonotomies increase DRG NKCC1 phosphorylation.4 NKCC1 KO mice have reduced Aβ-fiber mediated touch evoked hyperalgesia following intradermal capsaicin injections, a known method to induce allodynia.60 Intracolonic injection of capsaicin increases dorsal spinal NKCC1 phosphorylation within 10 minutes of injection and membrane mobilization 90-180 minutes after instillation. Total NKCC1 mRNA levels do not change.61 Intrathecal (IT) injections of BU have antinociceptive properties for hindpaw formalin injection models, a known method to induce acute pain.62 IT injections of BU also attenuates intracolonic capsaicin injection induced referred abdominal allodynia after its establishment.63 Recently, it was shown that IT BU injections reduces dorsal horn and nociceptive specific signaling after intraplantar capsaicin injections.64

Adult male Sprague-Dawley rats after cSCI show increased NKCC1 channel and decreased KCC2 channel expression 2-14 days post cSCI at the injury epicenter. Injured rats develop thermal hyperalgesia (TH) 21-42 days post cSCI. Administration of BU increases noxious thermal paw withdrawal latency time, signaling decreased TH. NKCC1 and KCC2 expression did not change in sham control animals in this experiment. This suggests the role of NKCC1 and KCC2 in the role of development and maintenance of cSCI induced NP.16 Inflammatory mediators induce phosphorylation of DRG NKCC1 channels and increases [Cl-]i within one hour, and increases NKCC1 expression and decreases KCC2 channel expression within three hours, in vitro.65

Hemisection spinal cord injury (SCI) decreases KCC2 expression in the dorsal horn that correlates with ≥ twelve-week mechanical allodynia. This type of injury also results in a positive shift in GABAi that changes prior inhibitory post-synpatic potentials to long lasting excitatory post-synaptic potentials in laminae I dorsal horn neurons.10 cSCI rats show a 84% reduction in ventral horn KCC2 channel expression 7-45 days after injury, and continuously decreased expression into 4-5 month post-injury chronic phases.66 Spinal cord KCC2 protein levels are decreased in rats with painful diabetic neuropathy.67 IT injections of anti-sense KCC2 oligodeoxynucleotides or a KCC2 channel blocker decreases mechanical and thermal nociceptive thresholds in injured and uninjured animals.68,69 Rat hindpaw formalin injection models show reduced KCC2 immunoreactivity in lamina I and II of L5, although total KCC2 mRNA is unchanged.70 Mice given subcutaneous injections of formalin show reduced KCC2 channel expression in the medullary horn that is associated with pain behaviors.69 Peripheral inflammation induced by hindpaw injections of complete Freund’s adjuvant reduces dorsal horn KCC2 channel expression and thermal nociceptive thresholds.71 Cuff-induced injuries of the rat sciatic nerve results in reduced expression of the KCC2 channel, and reverses GABA response polarity to excitatory in lamina I neurons, in vitro.68 In rat vagal motorneurons, in vivo axonotomies result in decreased expression of KCC2 mRNA. Subsequent accumulation of [Cl-]i is directly attributable to new GABA induced excitation.15

Role of GABA in Pain

GABA receptors are found in primary afferent terminals and interneurons in laminae I-IV in the spinal cord dorsal horn,72,73 which is the main site of Aδ- and C-fiber afferent termination and nociceptive signaling. GABA-ergic interneurons are important in spinal nociceptive processing and nociceptive attenuation.74,75 Elevation of [Cl-]i can lead to GABA-ergic hypersensitivity by reversing both ECl- and the normal inhibitory action of GABA.24 Lamina I GABA-ergic interneurons become more excitable with depolarizing membrane potentials, larger spike heights, increased firing frequencies and increased incidence of spontaneous plateau potentials after SCI.76 The GABA antagonist bicuculine alleviates formalin induced tactile allodynia in rats with painful diabetic neuropathy.67 Administration of complete Freund’s adjuvant into the rat hindpaw reverses spinal GABAa signaling. Muscimol (a GABAa receptor agonist) increases and gabazine (a GABAa antagonist) decreases nociceptive thresholds in naïve rats, where as after inflammation muscimol decreases and gabazine increases nociceptive thresholds.77 In vitro scraping injuries to hypothalamic neurons changes their electrophysiological properties: depolarizing chloride reversal potentials that result in GABA induced excitation.78

Future Perspectives

In summary, WNK1 phosphorylates SPAK and OSR1, which go onto phosphorylate NKCC1 and KCC2, activating and deactivating these channels, respectively. Subsequent accumulation of [Cl-]i could reverse GABA polarity in dorsal horn spinal interneurons. Altered WNK1 expression could be important in post-injury neuronal signaling.

SCI is a devastating79 and costly injury with an estimated 12,000 new cases reported within the US each year.80 Anywhere between 25.5-96.2% of people develop chronic pain after their injury.81-83 Neuropathic pain (NP) can occur from altered processing in the central nervous system in the absence of peripheral nerve damage.84 PAD and presynaptic inhibition could be modified by changes in WNK1 activity and/or expression, and subsequent changes in NKCC1 and KCC2 channel activity after cSCI to alter nociceptive sensory processing in the spinal cord. Altering these channels would change [Cl-]i and result in a larger potential shift when GABAa receptor channels open. This could lead to PAD changing to DRR and/or increased GABA activity of interneurons mediating PAD; ultimately leading to heightened excitability that would translate into hyperalgesia and allodynia (Fig. 1). Future electrophysiological studies could help understand post-injury changes in spinal circuitry.

An external file that holds a picture, illustration, etc.
Object name is ANS0972-7531-18-070-g001.jpg

Hypothetical role of WNK1 in pathologic spinal cord signaling. In normal spinal cord signaling tactile information is processed by Aβ-fibers, and a presynaptically linked GABA-ergic interneuron causes PAD of nociceptive pathways. However, an unknown mechanism such as injury will a. cause phosphorylation of WNK1 which, b. phosphorylates OSR1 and SPAK which, c. phosphorylates the NKCC1 and KCC2 channels, activating and deactivating these channels, respectively. This leads to [Cl-]i >ECl.-, d. reversed GABA signaling, and e. activation of otherwise silent nociceptive pathways and antidromically conducted DRR, leading to hyperalgesia or allodynia.

The hyperosmotic induced WNK1 and NKCC1 activation, and KCC2 deactivation previously reported in the renal system17,37,39,44 could be similar to the inflammatory response elicited in the nervous system after injury. Vasodilatation and subsequent invasion from neutrophils, monocytes, T and B lymphocytes; and cytokine secretion from astrocytes, microglial cells, endothelial cells and leukocytes85 could possibly increase extracellular osmolarity and activate WNK1, which has previously been described as a volume sensitive kinase.17,37,39,44 NKCC1 and KCC2 expression is increased and decreased, respectively, at a CNS injury center.16 NKCC1 phosphorylation stimulates peripheral nerve regrowth after axonotomy.59 Perhaps during injury GABA induced depolarizations, because of altered chloride homeostasis, are induced in an attempt to revert neurons back to a state of developmental flexibility needed for sprouting and retargeting.86 As a consequence, NP develops.

SLC12 channel phosphorylation precedes changed channel expression in nervous system injury models.65 And although NKCC1 phosphorylation has been shown to increase membrane mobilization,61 an exact role of WNK1 and increased SLC12 channel expression has not been described. Future studies directed at studying the consequences of altered WNK1 expression in the CNS will be important in understanding the various roles of this kinase.

Abbreviations

WNK1: with no lysine (K) kinase I; SPAK: STE20/SPS1-related proline/alanine-rich kinase; OSR1: oxidative stress-responsive kinase-1; NKCC1: Na+-K+-Cl--cotransporter I; KCC2: K+-Cl--cotransporter 2; GABA: γ-aminobutyric-acid; SCI: spinal cord injury; NP: neuropathic pain; DRG: dorsal root ganglion; BU: bumetanide; cSCI: contusion spinal cord injury; [Cl-]i: intracellular chloride; Cl-in/out: chloride current in/out; ECl-: chloride equilibrium potential; Vm: membrane potential; KO: knock-out; WT: wild-type; CNS: central nervous system; Pseudohypoaldosteronism type II: PHAII; HSAN2: hereditary sensory and autonomic neuropathy type 2; PAD: primary afferent depolarization; DRR: dorsal root reflexes; IT: intrathecal; TH: thermal hyperalgesia; GABAi: GABA induced current.

Acknowledgments

EMK would like to thank members of the Dr. Daniel K. Resnick and Dr. Gurwattan S. Miranpuri Spinal Cord Injury Research Lab at the University of Wisconsin School of Medicine and Public Health, Madison, WI, USA. They provided all materials, instruction and guidance necessary to prepare this manuscript.

Footnotes

The article complies with International Committee of Medical Journal Editor’s uniform requirements for the manuscripts.

Competing interests – None

Source of Funding – None

References

1. Payne JA, Xu JC, Haas M et al. Primary structure, functional expression, and chromosomal localization of the Bumetanide-sensitive Na-K-Cl cotransporter in human colon. J Biol Chem. 1995;270:17977–17985. [PubMed] [Google Scholar]
2. Yan Y, Dempsey RJ, Sun D. Na-K-Cl cotransporter in rat focal cerebral ischemia. J Cereb Blood Flow Metab. 2001;21:711–721. [PubMed] [Google Scholar]
3. Plotkin MD, Kaplan MR, Peterson LN et al. Expression of the Na-K-2Cl cotransporter BSC2 in the nervous system. Am J Physiol. 1997;272:C173–C183. [PubMed] [Google Scholar]
4. Delpire E, Austin TM. Kinase regulation of the Na-K-2Cl cotransporter in primary afferent neurons. J Physiol. 2010;588:3365–3373. [PMC free article] [PubMed] [Google Scholar]
5. Sung KW, Kirby M, McDonald MP et al. Abnormal gabaa receptor-mediated currents in dorsal root ganglion neurons isolated from Na-K-2Cl cotransporter null mice. J Neurosci. 2000;20:7531–7538. [PMC free article] [PubMed] [Google Scholar]
6. Kahle KT, Rinehart J, Ring A et al. WNK protein kinases modulate cellular Cl flux by altering the phosphorylation state of the Na-K-Cl and K-Cl cotransporters. Physiol. 2006;21:326–335. [PubMed] [Google Scholar]
7. Kaji DM. Effects of membrane potential on K-Cl transport in human erythrocytes. Am J Physiol Cell Physiol. 1993;68:C376–C382. [PubMed] [Google Scholar]
8. Payne JA, Stevenson TJ, Donaldson LF. Molecular characterization of a putative K-Cl cotransporter in rat brain - A neuronal-specific isoform. J Biol Chem. 1996;273:16245–16252. [PubMed] [Google Scholar]
9. Hubner CA, Stein V, Borgmeyer IH et al. Disruption of KCC2 reveals an essential role of K-Cl cotransport already in early synaptic inhibition. Neuron. 2001;30:515–524. [PubMed] [Google Scholar]
10. Lu Y, Zheng J, Xiong L et al. Spinal cord injury-induced attenuation of GABA-ergic inhibition in spinal dorsal horn circuits is associated with down-regulation of the chloride transporter KCC2 in rat. J Physiol. 2008;586:5701–5715. [PMC free article] [PubMed] [Google Scholar]
11. Gautam MS, John J, Rout J et al. Protective effects of graded doses of Gabapentin on Aminophylline-induced experimental status epilepticus in mice. Annals of Neurosciences. 2009;16(4):150–154. [Google Scholar]
12. Woo NS, Lu J, England R et al. Hyperexcitability and epilepsy associated with disruption of the mouse neuronal-specific K-Cl cotransporter gene. Hippocampus. 2002;12:258–268. [PubMed] [Google Scholar]
13. Wake H, Watanabe M, Moorhouse AJ et al. Early changes in KCC2 phosphorylation in response to neuronal stress result in functional downregulation. J Neurosci. 2007;27:1642–1650. [PMC free article] [PubMed] [Google Scholar]
14. Billups D, Attwell D. Control of intracellular chloride concentration and GABA response polarity in rat retinal ON bipolar cells. J Physiol. 2002;545:183–198. [PMC free article] [PubMed] [Google Scholar]
15. Nabekura J, Ueno T, Okabe A et al. Reduction of KCC2 expression and GABA receptor-mediated excitation after in vivo axonal injury. J Neurosci. 2002;22:4412–4417. [PMC free article] [PubMed] [Google Scholar]
16. Cramer SW, Baggot C, Cain J et al. The role of cation-dependent chloride transporters in NP following spinal cord injury. Mol Pain. 2008;4:36–44. [PMC free article] [PubMed] [Google Scholar]
17. Kahle KT, Rinehart J, de los Heros P et al. WNK3 modulates transport of Cl- in and out of cells: Implications for control of cell volume and neuronal excitability. Proc Natl Acad Sci. 2005;102:16783–16788. [PMC free article] [PubMed] [Google Scholar]
18. Jennings MI, Al-Rohil N. Kinetics of activation and inactivation of swelling-stimulated K/Cl transport the volume-sensitive parameter is the rate constant for inactivation. J Gen Physiol. 1990;95:1021–1040. [PMC free article] [PubMed] [Google Scholar]
19. Flatman PW. Regulation of Na-K-2Cl cotransport by phosphorylation and protein-protein interactions. Biochim Biophys Acta. 2002;1566:140–151. [PubMed] [Google Scholar]
20. Darman RB, Forbush B. A regulatory locus of phosphorylation in the N terminus of the Na-K-Cl cotransporter, NKCC1. J Biol Chem. 2002;277:37542–37550. [PubMed] [Google Scholar]
21. Shimura M, Akaike N, Harata N. Circadian rhythm in intracellular Cl- activity of acutely dissociated neurons of suprachiasmatic nucleus. Am J Physiol Cell Physiol. 2002;282:366–373. [PubMed] [Google Scholar]
22. Wagner S, Castel M, Gainer H et al. GABA in the mammalian suprachiasmatic nucleus and its role in diurnal rhythmicity. Nature. 1997;387:598–603. [PubMed] [Google Scholar]
23. Misgeld U, Deisz RA, Dodt HU et al. The role of chloride transport in postsynaptic inhibition of hippocampal neurons. Science. 1986;232:1413–1415. [PubMed] [Google Scholar]
24. Blaesse P, Airaksinen MS, Rivera C et al. Cation-chloride cotransporters and neuronal function. Neuron. 2009;61:820–838. [PubMed] [Google Scholar]
25. Rivera C, Hong L, Thomas-Crusells J et al. BDNF-induced TrkB activation down-regulated the K-Cl cotransporter KCC2 and impairs neuronal Cl- extrusion. J Cell Biol. 2002;159:747–752. [PMC free article] [PubMed] [Google Scholar]
26. Plotkin MD, Synder EY, Herbert SC et al. Expression of the Na-K-2Cl cotransporter is developmentally regulated in postnatal rat brains: A possible mechanism underlying GABA’s excitatory role in immature brain. J Neurobiol. 1997;33:781–795. [PubMed] [Google Scholar]
27. Ge S, Goh ELK, Sailor KA et al. GABA regulates synaptic integration of newly generated neurons in the adult brain. Nature. 2006;439:589–593. [PMC free article] [PubMed] [Google Scholar]
28. Sun D, Murali SG. Na-K-2Cl cotransporter in immature cortical neurons: A role in intracellular Cl- regulation. J Neurophysiol. 1999;81:1939–1948. [PubMed] [Google Scholar]
29. Ben-Ari Y. Excitatory actions of GABA during development: The nature of the nurture. Nature Rev Neurosci. 2002;3:728–739. [PubMed] [Google Scholar]
30. Rivera C, Voipio J, Payne JA et al. The K/Cl cotransporter KCC2 renders GABA hyperpolarizing during neuronal maturation. Nature. 1999;397:251–255. [PubMed] [Google Scholar]
31. Cherubini E, Gaiarsa JL, Ben-Ari Y. GABA: an excitatory transmitter in early postnatal life. Trends Neurosci. 1991;12:515–519. [PubMed] [Google Scholar]
32. Xu B, English JM, Wilsbacher JL et al. WNK1, a novel mammalian serine-threonine protein kinase lacking the catalytic lysine in subdomain II. J Biol Chem. 2002;275:16795–16801. [PubMed] [Google Scholar]
33. Delaloy C, Lu J, Houot AM et al. Multiple promoters in the WNK1 gene: One controls expression of kidney-specific kinase-defective isoform. Mol Cell Biol. 2003;23:9208–9220. [PMC free article] [PubMed] [Google Scholar]
34. Shekarabi M, Girard N, Riviere JP et al. Mutations in the nervous system – specific HSN2 exon on WNK1 cause hereditary sensory neuropathy type II. J Clin Invest. 2008;118:2496–2505. [PMC free article] [PubMed] [Google Scholar]
35. Wilson FH, Disee-Nicodeme S, Choate KA et al. Human hypertension caused by mutations in WNK kinases. Science. 2001;293:1107–1112. [PubMed] [Google Scholar]
36. Lenertz LY, Lee BH, Min X et al. Properties of WNK1 and implications for other family members. J Biol Chem. 2005;280:26653–26658. [PubMed] [Google Scholar]
37. Cain H, Baggott C, Tilghman JI et al. Recent Developments in the Study of Spinal Cord Injury and Neuropathic Pain. Annals of Neurosciences. 2007;14(4):96–107. [Google Scholar]
38. Vitari AC, Thastrup J, Rafiqi FH et al. Functional interactions of the SPAK/OSR1 kinases with their upstream activator WNK1 and downstream substrate NKCC1. Biochem J. 2006;397:223–231. [PMC free article] [PubMed] [Google Scholar]
39. Vitari AC, Deak M, Morrice NA et al. The WNK1 and WNK4 protein kinases that are mutated in Gordon’s hypertension syndrome phosphorylate and activate SPAK and OSR1 protein kinases. Biochem J. 2005;391:17–24. [PMC free article] [PubMed] [Google Scholar]
40. Moriguchi T, Urushlyama S, Hisamoto N et al. WNK1 regulates phosphorylation of cation-chloride-coupled cotransporters via the STE20-related kinases, SPAK and OSR1. J Biol Chem. 2005;280:42685–42693. [PubMed] [Google Scholar]
41. Delpire E, Gagnon KBE. SPAK and OSR1: STE20 kinases involved in the regulation of ion homoeostasis and volume control in the mammalian cells. Biochem J. 2008;409:321–331. [PubMed] [Google Scholar]
42. Gagnon KBE, England R, Delpire E. A single binding motif is required for SPAK activation of the Na-K-2Cl cotransporter. Cell Physiol Biochem. 2007;20:131–142. [PubMed] [Google Scholar]
43. Gagnon KBE, England R, Delpire E. Volume sensitivity of cation-Cl-cotransporters is modulated by the interaction of two kinases: STE20-related proline-alanine-rich kinase and WNK4. Am J Cell Physiol. 2006;290:134–142. [PubMed] [Google Scholar]
44. Ohta A, Yang SS, Rai T et al. Overexpression of human WNK1 increases paracellular chloride permeability and phosphorylation of claudin-4 in MDCKII cells. Biochem Biophys Res Comm. 2006;349:804–808. [PubMed] [Google Scholar]
45. Kahle KT, Ring AM, Lifton RP. Molecular physiology of the WNK kinases. Ann Rev Physiol. 2008;70:329–355. [PubMed] [Google Scholar]
46. Sun X, Gao L, Yu RK et al. Down-regulation of WNK1 protein kinase in neural progenitor cells suppresses cell proliferation and migration. J Neurochem. 2006;99:1114–1121. [PubMed] [Google Scholar]
47. Ushiro H, Tsutsumi T, Suzuki K et al. Molecular cloning and characterization of a novel STE20-related protein kinase enriched in neurons and transporting epithelia. Arch Biochem Biophys. 1998;355:233–240. [PubMed] [Google Scholar]
48. Miao N, Fung B, Sanchez R et al. Isolation and expression of PASK, a serine/threonine kinase, during rat embryonic development, with special emphasis on the pancreas. J Histochem Cytochem. 2000;48:1391–1400. [PubMed] [Google Scholar]
49. Geng Y, Hoke A, Delpire E. The STE20 kinases STE20-related proline-alaninerich kinase and oxidative-stress response 1 regulate NKCC1 function in sensory neurons. J Biol Chem. 2009;284:14020–14028. [PMC free article] [PubMed] [Google Scholar]
50. Arion D, Lewis DA. Altered expression of regulators of the cortical chloride transporters NKCC1 and KCC2 in schizophrenia. Arch Gen Psychiatry. 2011;68:21–31. [PMC free article] [PubMed] [Google Scholar]
51. Lundkvist GB, Kristensson K, Hill RH. Suprachiasmatic nucleus exhibits diurnal variations in spontaneous excitatory postsynaptic activity. J Biol Rhythms. 2002;17:40–51. [PubMed] [Google Scholar]
52. Loggia ML, Bushnell MC, Tetreault M et al. Carriers of recessive WNK1/HSN2 mutations for hereditary sensory and autonomic neuropathy type 2 (HSAN2) are more sensitive to thermal stimuli. J Neurosci. 2009;29:2162–2166. [PMC free article] [PubMed] [Google Scholar]
53. Delaloy C, Elvira-Matelot E, Clemessy M et al. Deletion of WNK1 first intron in misregulation of both isoforms in renal extrarenal tissues. Hypertension. 2010;52:1149–1154. [PubMed] [Google Scholar]
54. Woolf CJ, Salter MW. Neuronal plasticity: increasing the gain in pain. Science. 2000;288:1765–1768. [PubMed] [Google Scholar]
55. Cervero F, Laird JMA. Mechanisms of touch-evoked pain (allodynia): a new model. Pain. 1996;68:13–23. [PubMed] [Google Scholar]
56. Cervero F, Laird JMA, Garcia-Nicas E. Secondary hyperalgesia and presynaptic inhibition: an update. Eur J Pain. 2003;7:345–351. [PubMed] [Google Scholar]
57. Curtis DR, Lodge D. The depolarisation of feline ventral horn group Ia spinal afferent terminations by GABA. Exp Brain Res. 1982;46:215–233. [PubMed] [Google Scholar]
58. Willis WD. Dorsal root potentials and dorsal root reflexes: a double-edged sword. Exp Brain Res. 1999;124:395–421. [PubMed] [Google Scholar]
59. Pieraut S, Laurent-Matha V, Sar C et al. NKCC1 phosphorylation stimulates neurite growth of injured adult sensory neurons. J Neurosci. 2007;27:6751–6759. [PMC free article] [PubMed] [Google Scholar]
60. Laird JMA, Garcia-Nicas E, Delpire EJ et al. Presynaptic inhibition and spinal pain processing in mice: a possible role of the NKCC1 cation-chloride co-transporter in hyperalgesia. Neurosci Lett. 2004;361:200–203. [PubMed] [Google Scholar]
61. Galan A, Cervero F. Painful stimuli induce in vivo phosphorylation and membrane mobilization of the mouse spinal cord NKCC1 co-transporter. Neurosci. 2005;133:245–252. [PubMed] [Google Scholar]
62. Granados-Soto V, Arguelles CF, Alvarez-Leefmans FJ. Peripheral and central antinociceptive action of Na-K-2Cl cotransporter blockers on formalin-induced nociception in rats. Pain. 2005;114:231–238. [PubMed] [Google Scholar]
63. Pitcher MH, Price TJ, Entrena JM et al. Spinal NKCC1 blockade inhibits TRPV1-dependent referred allodynia. Mol Pain. 2007;3:17–25. [PMC free article] [PubMed] [Google Scholar]
64. Pitcher MH, Cervero F. Role of the NKCC1 co-transporter n in sensitization of spinal nociceptive neurons. Pain. 2010;151:756–762. [PubMed] [Google Scholar]
65. Funk K, Woitecki A, Franjic-Wurtz C et al. Modulation of chloride homeostasis by inflammatory mediators in dorsal root ganglion neurons. Mol Pain. 2008;4(32) [PMC free article] [PubMed] [Google Scholar]
66. Boulenguez P, Liabeuf S, Bos R et al. Down-regulation of the potassium-chloride cotransporter KCC2 contributes to spasticity after spinal cord injury. Nature Med. 2010;16:302–308. [PubMed] [Google Scholar]
67. Jolivalt CG, Lee CA, Ramos KM et al. Allodynia and hyperalgesia in diabetic rats are mediated by GABA and depletion of spinal potassium-chloride co-transporters. Pain. 2008;140:48–57. [PMC free article] [PubMed] [Google Scholar]
68. Coull JAM, Boudreau D, Bachand K et al. Trans-synaptic shift in anion gradient in spinal lamina I neurons as a mechanism of NP. Nature. 2003;424:938–942. [PubMed] [Google Scholar]
69. Wu LA, Huang J, Wang W et al. Down-regulation of the K-Cl co-transporter 2 in mouse medullary dorsal horn contributes to the formalin-induced inflammatory orofacial pain. Neurosci Lett. 2009;457:36–40. [PubMed] [Google Scholar]
70. Nomura H, Sakai A, Nagano M et al. Expression changes of cation chloride cotransporters in the rat spinal cord following intraplantar formalin. Neurosci Res. 2006;56:435–440. [PubMed] [Google Scholar]
71. Zhang W, Liu LY, Xu TL. Reduced Potassium-Chloride co-transporter expression in spinal cord dorsal horn neurons contributes to inflammatory pain hypersensitivity in rats. Neurosci. 2008;152:502–510. [PubMed] [Google Scholar]
72. Bowery NG, Hudson AL, Price GW. GABAa and GABAb receptor site distribution in the rat central nervous system. Neurosci. 1987;20:365–383. [PubMed] [Google Scholar]
73. Todd AJ, McKenzie J. GABA-immunoreactive neurons in the dorsal horn of the rat spinal cord. Neurosci. 1989;31:799–806. [PubMed] [Google Scholar]
74. Price TJ, Cervero F, de Koninck Y. Role of Cation-Chloride-Cotransporters (CCC) in pain and hyperalgesia. Curr Top Med Chem. 2005;5:547–555. [PMC free article] [PubMed] [Google Scholar]
75. Price TJ, Cervero F, Gold MS et al. Chloride regulation in the pain pathway. Brain Res Rev. 2009;60:149–170. [PMC free article] [PubMed] [Google Scholar]
76. Dougherty KJ, Hochman S. Spinal cord injury causes plasticity in a subpopulation of lamina I GABA-ergic interneurons. J Neurophysiol. 2008;100:212–223. [PMC free article] [PubMed] [Google Scholar]
77. Anseloni VCZ, Gold MS. Inflammation-induced shift in the valence of spinal GABA-A receptor-mediated modulation of Nociception in the adult rat. J Pain. 2008;9:732–738. [PMC free article] [PubMed] [Google Scholar]
78. van den Pol AN, Obrietan K, Chen G. Excitatory actions of GABA after neuronal trauma. J Neurosci. 1996;16:4283–4292. [PMC free article] [PubMed] [Google Scholar]
79. Siddall PJ, McCelland JM, Rutkowski SB et al. A longitudinal study of the prevalence and characteristics of pain in the first 5 years following spinal cord injury. Pain. 2003;103:249–257. [PubMed] [Google Scholar]
80. Spinal Cord Injury Facts and Figures at a Glance – February 2010. Retrieved; from. National Spinal Cord Injury Statistic Center (NSCISC). Jul 9, 2010. https://www.nscisc.uab.edu/public_content/pdf/Facts%20and%20Figures%20at%20a%20Glance%202010.pdf.
81. Lundqvist C, Siosteen A, Blomstrand C et al. Spinal cord injuries - clinical, functional, and emotional status. Spine. 1991;16:78–83. [PubMed] [Google Scholar]
82. Raissi GR, Mokhtari A, Mansouri K. Reports from spinal cord injury patients eight months after the 2003 earthquake in Bam, Iran. Am J Phys Med Rehabil. 2007;86:912–917. [PubMed] [Google Scholar]
83. Dijkers M, Bryce TB, Zanca J. Prevalence of chronic pain after traumatic spinal cord injury: A systematic review. J Rehabil Res Dev. 2009;6:13–30. [PubMed] [Google Scholar]
84. Gracely RH, Lynch SA, Bennett GJ. Painful neuropathy: altered central processing maintained dynamically by peripheral input. Pain. 1992;51:175–194. [PubMed] [Google Scholar]
85. Ankeny DP, Popovich PG. Mechanisms and implications of adaptive immune responses after traumatic spinal cord injury. Neuroscience. 2009;158:1112–1121. [PMC free article] [PubMed] [Google Scholar]
86. Payne JA, Rivera C, Voipio J et al. Cation-chloride co-transporters in neuronal communication, development and trauma. Trends Neurosci. 2003;26:199–206. [PubMed] [Google Scholar]

Articles from Annals of Neurosciences are provided here courtesy of SAGE Publications

-