Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2012 Mar 1.
Published in final edited form as: Expert Rev Clin Pharmacol. 2011 May 1;4(3):379–388. doi: 10.1586/ecp.11.17

Targeting N-methyl-D-aspartate receptors for treatment of neuropathic pain

Hong-Yi Zhou 1, Shao-Rui Chen 1, Hui-Lin Pan 1,2,
PMCID: PMC3113704  NIHMSID: NIHMS295692  PMID: 21686074

Abstract

Neuropathic pain remains a major clinical problem and a therapeutic challenge because existing analgesics are often ineffective and can cause serious side effects. Increased N-methyl-D-aspartate receptor (NMDAR) activity contributes to central sensitization in certain types of neuropathic pain. NMDAR antagonists can reduce hyperalgesia and allodynia in animal models of neuropathic pain induced by nerve injury and diabetic neuropathy. Clinically used NMDAR antagonists, such as ketamine and dextromethorphan, are generally effective in patients with neuropathic pain, such as complex regional pain syndrome and painful diabetic neuropathy. However, patients with postherpetic neuralgia respond poorly to NMDAR antagonists. Recent studies on identifying NMDAR-interacting proteins and molecular mechanisms of increased NMDAR activity in neuropathic pain could facilitate the development of new drugs to attenuate abnormal NMDAR activity with minimal impairment of the physiological function of NMDARs. Combining NMDAR antagonists with other analgesics could also lead to better management of neuropathic pain without causing serious side effects.

Keywords: complex regional pain syndrome, diabetic neuropathy, neuropathic pain, NMDA receptors, spinal cord, synaptic plasticity, synaptic transmission


Chronic pain, such as neuropathic pain caused by peripheral nerve injury and diabetic neuropathy, can result in prolonged suffering and reduced quality of life. Pain can occur spontaneously or as a result of exposure to mildly painful stimuli (hyperalgesia) or stimuli not normally perceived as painful (allodynia). However, the treatment of chronic pain remains a major challenge, with only approximately 50% of patients receiving adequate pain relief [1].

The glutamate N-methyl-D-aspartate receptors (NMDARs), especially those located in the dorsal horn of the spinal cord, are critically involved in nociceptive transmission and synaptic plasticity and have long been considered a target for the treatment of neuropathic pain. However, despite numerous preclinical studies demonstrating the efficacy of NMDAR antagonists in treating neuropathic pain, clinical studies suggest that NMDAR antagonists have limited effects in a subpopulation of patients with chronic neuropathic pain. Here, we review the current literature about the function and regulation of NMDARs and the therapeutic effects of NMDAR antagonists on chronic neuropathic pain.

Structure, function & distribution of NMDARs

N-methyl-D-aspartate receptors are heteromeric protein complexes, and three families of NMDAR subunits have been identified: NR1, NR2 and NR3. There are eight different NR1 (GluN1) subunits, four different NR2 subunits (NR2A-2D and GluN2A-2D) and two NR3 subunits (NR3A and NR3B, and GluN3A and GluN3B). While the eight different NR1 subunits are generated by alternative splicing from a single gene [2], the NR2 and NR3 subunits are encoded by six separate genes [37]. NMDARs are tetramers that typically incorporate two NR1 subunits plus two NR2 subunits and, in some cases, include an NR3 subunit [8]. The subunit composition determines the pharmacological and physiological properties of the NMDARs [4,9]. NR1/NR2A channels display much faster inactivation than do NR1/NR2B, NR1/NR2C and NR1/NR2D subunits [9].

The NR1 subunits are ubiquitously distributed in all neurons and at all developmental stages in the CNS [10]. NR1/NR2A subunits become predominant as synapses stabilize and mature [11]. NR2A is gradually increased in the adult CNS but is absent in the embryonic CNS [10,12]. In early developmental stages, NR2B is abundant, but it declines with age and becomes more restricted to the forebrain, including the cortex, hippocampus, striatum, thalamus and olfactory bulb [10,12]. Electrophysiological data suggest that NR2A-containing NMDARs tend to be localized synaptically, whereas NR2B-containing receptors tend to be localized perisynaptically or extrasynaptically [1315]. In the spinal cord, the predominant subunits of NMDARs include NR2A at synaptic sites and NR2B at extrasynaptic sites [16]. NR2C is absent at early developmental stages but becomes progressively predominant in the cerebellum. NR2D is mainly found in the thalamus, hippocampus and brainstem [12,17]. Of note, NR1, NR2A, NR2B and NR2D have been detected in the superficial dorsal horn of the spinal cord [16,1820]. NR1 subunits are ubiquitously distributed in all laminae of the spinal cord, but the distribution of NR2B subunits appears to be restricted to the superficial dorsal horn of the spinal cord [19]. In addition, the NR1 subunits are present in the cell bodies and central terminals of both small- and large-diameter dorsal root ganglia neurons [2123]. Dorsal root ganglia neurons also contain NR2B, NR2C and NR2D, but not NR2A [23,24].

In the three families of NMDAR subunits, NR1 binds to glycine, NR2 binds to glutamate and NR3 is found to bind to glycine. Thus, NMDARs composed of NR1 and NR3 subunits only require glycine for activation. The recent discovery of the functionally different NR3 subunits has caused many of NMDARs’ known features to be reassessed [6,8,25,26]. NR3A can coassemble with NR1 and NR2A to form a receptor complex with distinct single-channel properties and reduced relative calcium permeability [26]. Interestingly, cotransfection of NR1, NR2A and NR3B into HEK293 cells results in significantly lower amplitude and calcium-permeability of the whole-cell currents than that permitted by NR1/NR2A receptors [7,25]. In addition, increasing the proportion of NR3B cDNA relative to NR1 and NR2A cDNA significantly reduces the NMDAR current amplitude [7]. A reduced shift in reversal potential, induced by switching the extracellular Ca2+ concentration from 1 to 20 mM, can be detected in NR3B-overexpressed hippocampal neurons (16.1 mV) compared with native neurons (27.2 mV) [27]. Thus, NR3 subunits seem to negatively interfere with NMDAR activity. NR3A expression is relatively high at younger ages and then decreases with age [28], whereas NR3B expression gradually increases with age [29]. High levels of NR3A are present in the spinal cord, thalamus, hypothalamus, brainstem, CA1 of the hippocampus, amygdala and certain parts of the cortex [28]. NR3B is distributed in the forebrain, cerebellum and lumbar spinal cord [30]. Owing to the lack of selective agonists and antagonists, the function of NR3-containing NMDARs in synaptic transmission is not well understood.

Regulation of NMDAR function by subunit composition & phosphorylation

The unique feature of the NMDAR is its voltage and ligand dependence [31,32]. NR3A subunits have a larger influence on the voltage dependence of NMDARs [33]. The NMDAR is permeable to monovalent cations, including Na+ and K+, and divalent cations, most notably Ca2+. However, the NMDARs containing NR2C and NR2D are less sensitive to Mg2+ blocking than are NMDARs containing NR2A or NR2B [34]. The NMDAR’s voltage dependence follows directly from channel blocking by submillimolar concentrations of extracellular Mg2+, rather than from the voltage dependence of conformational changes [31,32]. At the resting membrane potential, Mg2+ largely blocks ion flow through the NMDAR channel. When the membrane is depolarized, Mg2+ is expelled from the channel, allowing for greatly enhanced passage of ions. The current–voltage relationship of typical NMDAR currents has a negative slope conductance [31,32]. In spinal dorsal horn neurons, the current–voltage curve shows the typical J shape of NMDAR-mediated currents [35]. Therefore, both depolarization of the postsynaptic neuron and presynaptic release of glutamate and glycine are required for maximum current flow through the NMDAR channel. Although glycine is considered a co-agonist for NMDARs, increased glycine release does not potentiate NMDAR currents in dorsal horn neurons [36]. Thus, an increase in synaptic glycine release probably activates inhibitory glycine receptors, rather than NMDARs, in the spinal dorsal horn.

The response of ionotropic glutamate receptors to agonists is usually potentiated after phosphorylation. NMDARs can be phosphorylated by protein kinase A (PKA) [3739], PKC [3840], casein kinase 2 (CK2) [41,42], Ca2+/calmodulin-dependent protein kinase II (CAMKII) [43,44] and tyrosine kinases Src [45] and Fyn [4648]. Phosphorylation by PKC increases the opening probability and decreases the affinity of NMDARs for extracellular Mg2+ [40]. Activation of PKC, possibly via upstream activation of group I metabotropic glutamate receptors, facilitates upregulation of NMDAR activity and enhances long-term potentiation (LTP) [49]. Ser-890, Ser-896 and Thr-879 in the C1 cassette of the NR1 subunit have been identified as PKC phosphorylation sites [38]. However, PKC-induced phosphorylation of NR1 at Ser-890 can inhibit clustering of NR1 [38], which might explain why stimulation of PKC suppresses NMDAR currents [50]. Although PKC directly phosphorylates the NR1 subunit [38,39,51], the PKC–cell adhesion kinase β/proline-rich tyrosine kinase 2–Src signaling cascade can indirectly upregulate NMDAR function [52]. In addition, PKC can phosphorylate Ser-1291 and Ser-1312 of NR2A [53] and Ser-1303 and Ser-1323 of NR2B [54] to potentiate currents through NR2A- or NR2B-containing NMDA channels.

Protein kinase A phosphorylates NR1 at Ser-879 in the C1 cassette of NMDARs [38]. In hippocampal neurons, PKA activation potentiates NMDAR activity indirectly by inhibiting calcineurin [37]. Activation of PKA is sufficient to induce synaptic clustering of NR1 [55]. By contrast, phosphorylation of NR1 by PKA can antagonize its interaction with spectrin and may initiate the morphological changes in dendrites associated with synaptic activity and plasticity [47]. CaMKII binds directly to NR1 and NR2B subunits, and activation of CaMKII by stimulation of NMDARs in the hippocampal slices increases their association [43]. Ser-1303 of NR2B and perhaps the homologous serine in NR2A are phosphorylation sites for CAMKII in hippocampal neurons [56]. Switching from NR2B-containing NMDARs (with high affinity for CaMKII) to NR2A-containing NMDARs (with low affinity for CaMKII) dramatically reduces LTP [44].

Activation of tyrosine kinases increases NMDAR-mediated responses in neurons [5759]. Approximately 2% of NR2A and 4% of NR2B, but not NR1, subunits are phosphorylated by tyrosine in synaptic plasma membranes [60]. Within the NMDAR complex, the nonreceptor tyrosine kinase Src is involved in the control of NMDAR activity through a cascade of intracellular signaling [45]. Thus, Src is considered to be an endogenous kinase that regulates NMDARs [45]. Tyr-1292, Tyr-Y1325 and Tyr-Y1387 of NR2A have been identified as the major Src-mediated phosphorylation sites [61]. NR2B is another major tyrosine-phosphorylated NMDAR subunit in the postsynaptic density [62], and its phosphorylation level is enhanced after LTP and by brain-derived neurotrophic factor treatments [63,64]. Seven specific tyrosine residues in the C-terminal cytoplasmic region of the NR2B subunit are phosphorylated by Fyn in vitro [65]. Of these seven residuals, three tyrosine phosphorylation sites (Tyr-1252, Tyr-1336 and Tyr-1472) are phosphorylated when active Fyn is coexpressed in the cell line [65]. Phosphorylation of NR2A by Src and Fyn kinases can potentiate NMDAR function [46]. Because Src and Fyn kinases have no effect on a receptor consisting of NR1 and a C-terminal truncation mutant of NR2A, the tyrosine phosphorylation sites may be targeted in the C-terminal domain of NR2A [46]. The potentiation of NR1/NR2B and NR1/NR2D by Src is much less than that of NR1/NR2A receptors [66].

In the hippocampus, the protein kinase CK2 seems to selectively potentiate the function of NMDARs but not that of α-amino-3-hydroxy-5-methylisoxazole-4-propionic acid (AMPA) receptors [41]. High-frequency synaptic activity induces a transient increase in CK2 activity during induction of LTP in an NMDAR-dependent manner [67]. CK2 selectively enhances synaptic NMDARs and induces LTP in the hippocampus [68]. It has been suggested that CK2 may increase NMDAR function through indirect mechanisms, possibly involving postsynaptic density (PSD)-95/synapse-associated protein (SAP)90, because the NR2A/B and NR1 subunits are not CK2 substrates [69]. However, others have demonstrated that CK2 phosphorylates Ser-1480 of the NR2B subunit within its PDZ binding domain [42]. Phosphorylation of Ser-1480 disrupts the interaction of NR2B with the PDZ domains of PSD-95 and SAP102 and decreases surface NR2B expression in cortical neurons [42]. Results from a recent study suggest that during neuronal development, increased CK2 activity phosphorylates NR2B Ser-1480 to drive NR2B endocytosis and remove NR2B from synapses, causing an increase in synaptic NR2A expression in cortical neurons [70]. Thus, CK2 may be important in determining the NR2 subunit composition of synaptic NMDARs.

Conversely, NMDAR activity can be inhibited by serine and threonine protein phosphatases. Phosphatases 1 and 2A reduce the opening probability of NMDARs in cultured hippocampal neurons [71]. Calcineurin also shortens the opening time of NMDARs in acutely dissociated dentate gyrus granule cells [72]. CK2 and calcineurin appear to have opposite effects on NMDAR channel gating [41]. Endogenous tyrosine phosphatases may also regulate the probability of NMDAR opening in spinal dorsal horn neurons [73].

Role of NMDARs in nociceptive transmission in neuropathic pain

Neuropathic pain caused by diabetic neuropathy and nerve injury is associated with increased glutamate release from primary afferent terminals and stimulation of AMPA receptors and metabotropic glutamate receptors, especially mGluR5, in the spinal cord [7477]. Although glutamate mainly acts on postsynaptic receptors to mediate excitatory neurotransmission, presynaptic NMDARs can increase the release of neurotransmitters such as substance P from the primary afferent terminals in the spinal dorsal horn [78]. Furthermore, presynaptic NMDARs mediate increased glutamate release from the primary afferent terminals and LTP triggered by μ-opioid receptor stimulation [79]. On the other hand, it has been reported that activation of presynaptic NMDARs inhibits electrically evoked glutamate release from the primary afferent terminals [80]. It is possible that stimulation of presynaptic NMDARs can indirectly reduce neurotransmitter release through the inhibition of voltage-activated Ca2+ channels because of Ca2+ influx and calcineurin stimulation [81].

The pathological role of NMDARs and the therapeutic effects of NMDAR antagonists on neuropathic pain have been studied primarily using animal models of peripheral nerve injury. Bath application of NMDA induces a greater increase in the whole-cell currents and calcium influx in spinal lamina II neurons in nerve-ligated rats than in control rats [82]. However, an increase in NMDAR activity in the spinal cord in diabetic neuropathy and postherpetic neuralgia has not yet been specifically documented. By directly recording NMDAR currents of dorsal horn neurons triggered by primary afferent stimulation, our preliminary data suggest that the activity of NMDARs in the spinal cord is increased in rats subjected to peripheral nerve ligation injury but not in the rat models of painful diabetic neuropathy and postherpetic neuralgia [83]. Partial ligation of the sciatic nerve significantly increases the phosphorylated proportion of the NR1 subunit of NMDARs in the dorsal horn [84]. Furthermore, PKA may increase NR1 subunit phosphorylation in the spinal cord in rats subjected to ligation of the L5 spinal nerve [85]. However, the total protein levels of NR1 and NR2A–2D subunits in the spinal cord remain unchanged after nerve injury [84,85]. Thus, increased NMDAR phosphorylation may be critical for central sensitization induced by nerve injury. In neuropathic pain induced by transection of the L5 spinal nerve, Fyn kinase phosphorylates the NR2B subunit at Tyr1472, and neuropathic pain is reduced in Fyn-knockout mice [48]. Furthermore, peripheral nerve injury can increase tyrosine phosphorylation of NR2B in the spinal cord, with no effect on the total NR2B protein level [86].

Wind-up is a frequency-dependent increase in the excitability of spinal cord neurons evoked by electrical stimulation of C-fiber primary afferent nerves. Wind-up has been interpreted as a system for the amplification in the spinal cord of the nociceptive message that arrives from peripheral nociceptors [87]. Blocking NMDARs with ketamine or (2R)-amino-5-phosphonopentanoate can largely inhibit the wind-up phenomenon of spinal dorsal horn neurons [88,89]. Blocking NMDARs reduces the hypersensitivity of spinal dorsal horn neurons in neuropathic pain models. The NMDAR antagonists ketamine, memantine and MK-801 potently reduce evoked responses of dorsal horn neurons in spinal nerve-ligated rats [90]. In addition, the NR2B subunit-specific antagonist ifenprodil reduces the amplitude of NMDAR currents in nerve-ligated mice [91]. Administration of Ro 25-6981, another selective NR2B antagonist, to the spinal cord not only decreases the C-fiber responses of dorsal horn neurons in both normal and spinal nerve-injured rats but also significantly blocks the LTP induced by high-frequency stimulation [92]. It has been reported that nerve injury decreases calcineurin levels in the spinal cord [93], suggesting that decreased calcineurin levels may contribute to the increase in NMDAR phosphorylation that accompanies neuropathic pain. In addition, nerve injury-induced increases in NMDAR activity in the spinal cord could result from the subunit switch from NR2B to NR2A, but this possibility has not yet been carefully investigated.

Effects of NMDAR antagonists in animal models of neuropathic pain

Animal studies of the antinociceptive effects of NMDAR antagonists have been conducted almost exclusively using rodents subjected to different types of peripheral nerve injuries. Various noncompetitive NMDAR antagonists (i.e., MK-801, ketamine, memantine and dextrorphan) decrease the development of allodynia and hyperalgesia following constrictive injury of the sciatic nerve [9497] and spinal nerve ligation [90,98,99]. Moreover, intrathecal application of amino-5-phosphonopentanoate, a competitive NMDAR antagonist, reduces mechanical allodynia caused by spinal cord injury [100]. Parenteral and oral administration of norketamine, a primary metabolite of ketamine, has been reported to alleviate mechanical and thermal hyperalgesia, but not tactile allodynia, in the sciatic nerve injury model without significant side effects [101]. Compared with ketamine and MK-801, the low-affinity NMDAR blockers memantine and neramexane display faster unblocking kinetics and greater voltage dependency. Thus, these agents may be able to inhibit the ongoing NMDAR activation presumed to occur in neuropathic pain with a lower incidence of side effects. Chronic administration of the noncompetitive NMDAR antagonists neramexane and memantine for 2 weeks produces persistent antinociceptive effects on mechanical hyperalgesia and allodynia in a rat model of diabetic neuropathic pain [102].

Several studies have demonstrated that selective glycine-site NMDAR antagonists such as L-701,324, MRZ2/576 and 5,7-dichlorokinurenic acid (5,7-DCK) reverse allodynia induced by sciatic and spinal nerve injuries [103,104]. Another glycine-site NMDAR antagonist, GV196771, also appears to be effective in the animal model of sciatic nerve injury [105,106]. Pyridazinoquinolinetriones are new glycine-site NMDAR antagonists with better aqueous solubility and oral bioavailability, and they can reduce neuropathic pain induced by sciatic nerve injury in rats [107].

Intrathecal application of the NR2B antagonists ifenprodil and Ro 25-6981 has been reported to dose-dependently reduce allodynia caused by spinal nerve ligation [92]. Furthermore, intrathecal administration of ifenprodil attenuates thermal hyperalgesia and mechanical allodynia induced by chronic compression of the dorsal root ganglia and in a model of bone cancer pain [108,109]. However, we found that intrathecal injection of ifenprodil produced no effect on tactile allodynia induced by spinal nerve ligation in rats [Chen S-R et al., Unpublished Data]. Systematic administration of the NR2B antagonist CP-101,606 or Ro 25-6981 decreases neuropathic pain in the sciatic nerve injury model with fewer side effects than existing NMDAR antagonists [19]. Because intrathecal injection of CP-101,606 does not inhibit allodynia caused by sciatic nerve injury [110], CP-101,606 may exert its antinociceptive effect at the supraspinal level. Thus, NR2B at the spinal level does not appear to be important for the maintenance of neuropathic pain induced by nerve injury.

Because of the essential role of NMDARs in many physiological functions, the ideal approach for treating chronic pain would be to preferentially inhibit enhanced NMDARs but not alter basal NMDAR channel activity. Src is anchored within the NMDAR complex through NADH dehydrogenase subunit 2, an adaptor protein [111]. Thus, blocking the interaction between the Src unique domain and NADH dehydrogenase subunit 2 can release Src from the NMDAR complex, separating the enzyme and substrate, thereby inhibiting Src-mediated upregulation of NMDAR activity [111]. Systemic or intrathecal injection of a synthetic peptide consisting of amino acids 40–49 of Src fused to the protein transduction domain of the HIV Tat protein can reverse tactile and cold hypersensitivity induced by sciatic nerve injury in rats [86]. In our preliminary study, we found that inhibition of CK2 can also normalize NMDAR activity in the spinal cord and attenuate tactile allodynia induced by spinal nerve ligation without altering NMDAR currents and nociception in normal control rats [83].

Clinical studies on the efficacy of NMDAR antagonists in neuropathic pain

Intravenous infusion of ketamine at anesthetic and subanesthetic doses produces significant pain relief in patients suffering from refractory complex regional pain syndrome [112,113]. However, ketamine infusion does not prevent chronic neuropathic pain caused by thoracotomy in patients [114]. Intranasal administration of low-dose ketamine also decreases pain scores in patients with neuropathic pain of various origins [115]. In addition, high doses of dextromethorphan show modest effects on diabetic neuropathic pain but have no effect on postherpetic neuralgia [116]. Intravenous infusion of another NMDAR antagonist, amantadine, reduces the intensity of ongoing postsurgical neuropathic pain in cancer patients [117]. However, oral amantadine is not clinically useful because of frequent intolerable side effects and insufficient reduction of neuropathic pain [118]. Systemic administration of high-affinity NMDAR antagonists, such as ketamine, can produce various adverse effects in patients, such as hallucinations, drowsiness, restlessness, dissociation, vivid dreams, and memory and motor deficits. However, when ketamine infusion is carefully controlled to target certain plasma levels, it can be a safe and effective treatment for patients with neuropathic pain. Several clinical studies have demonstrated that intravenous ketamine infusion effectively reduces pain scores in patients with complex regional pain syndrome [112,119121]. Oral administration of dextromethorphan also can reduce pain scores in patients with painful diabetic neuropathy, but it is not effective in patients with postherpetic neuralgia [116].

Although some new NMDAR antagonists have been tested clinically, they are no more effective than ketamine in reducing neuropathic pain. Memantine generally has a low incidence of CNS side effects because it exhibits fast blocking/unblocking kinetics and relatively strong voltage dependence at the NMDARs. Treatment with memantine significantly reduces pain scores in patients with complex regional pain syndrome caused by a traumatic injury to an upper extremity [122]. However, other studies have shown that despite its intriguing efficacy in animal models, memantine seems to have little effects on patients with painful diabetic neuropathy and postherpetic neuralgia [123], chronic limb pain [124] and sensory neuropathy associated with immunodeficiency virus infection [125].

Oral administration of a glycine-site NMDAR antagonist, GV196771, is not effective in reducing the intensity of evoked pain in patients with neuropathic pain (including diabetic neuropathy, postherpetic neuralgia and complex regional pain syndrome), although GV196771 does reduce the areas of static and dynamic mechanical allodynia [126]. The specificity and clinical efficacy of glycine-site NMDAR antagonists in neuropathic pain remain to be determined. In addition, a recent Phase II randomized, double-blind, placebo-controlled trial showed that an NR2B antagonist, radiprodil (RGH-896), has no significant effect on the mean daily pain scores of patients with diabetic neuropathy.

When combined with other classes of analgesics, NMDAR antagonists could have the potential for increased efficacy and reduced adverse effects. A double-blind study suggests that intravenous infusion of a low dose of ketamine plus oral gabapentin is safe and efficacious in reducing pain scores in patients with neuropathic pain secondary to spinal cord injury [127]. Combined treatments with intravenous ketamine and opioids also show improved efficacy in neuropathic pain in cancer patients [128]. In a Phase III clinical trial, zenvia, a combination of dextromethorphan and quinidine (a clinically used drug that can block Na+ and K+ channels and slow down the metabolism of dextromethorphan), lowers pain ratings more than placebo does in patients with diabetic neuropathic pain.

Expert commentary

N-methyl-D-aspartate receptors have been an attractive target for treatment of chronic neuropathic pain for two decades. Ketamine and dextromethorphan remain the most studied NMDAR antagonists in the clinical setting for the treatment of patients with neuropathic pain. NMDAR antagonists are very effective in reducing hyperalgesia and allodynia caused by nerve injury in animal models. However, clinical studies suggest that the therapeutic effects of this class of drugs are mostly limited to patients with complex regional pain syndrome and painful diabetic neuropathy. Further studies are warranted to determine whether increased NMDAR activity in the spinal cord after nerve injury results from direct phosphorylation of NR1 or a switch from NR2B to NR2A on the plasma membrane. On the basis of our current knowledge, targeting NR2B for neuropathic pain treatment does not seem to be effective or well justified. It would be interesting to determine whether overexpressing NR3A and/or NR3B at the spinal level could impair NMDAR function and reduce neuropathic pain.

It should be recognized that neuropathic pain is not a homogeneous disease condition, and different mechanisms may be involved in the neuropathic pain caused by traumatic nerve injury, diabetic neuropathy and viral infections. It is striking that most preclinical studies showing the importance of NMDARs in central sensitization, and the effects of NMDAR antagonists on neuropathic pain are conducted almost exclusively using ligation or transection of peripheral nerves in animal models. Increased NMDAR activity in the spinal cord has only been shown in traumatic nerve injury models, but this increased NMDAR activity has not been demonstrated directly in other neuropathic pain conditions such as painful diabetic neuropathy and postherpetic neuralgia. Furthermore, while spontaneous pain (i.e., pain scores reported by patients) is evaluated in most clinical studies, evoked pain or pain hypersensitivity (e.g., hyperalgesia and allodynia) is only measured in preclinical studies. These differences in pain assessments may account for the differences in the efficacy of NMDAR antagonists in neuropathic pain between clinical and animal studies. The possibility of a placebo effect should also be taken into account when interpreting the observed effects in patients.

N-methyl-D-aspartate receptor channel blockers, such as ketamine often produce several adverse effects due to their interference with normal NMDAR function in the CNS. An ideal treatment for neuropathic pain would be to reduce the increased activity of NMDARs while maintaining their physiological function. Identifying additional mechanisms of increased NMDAR activity after nerve injury could lead to the development of improved treatments for neuropathic pain. For example, inhibiting the NMDAR phosphorylation sites without blocking the channels themselves could improve the therapeutic window. In addition, targeting the interacting proteins and kinases that regulate NMDAR function at the spinal level, such as disrupting the Src-NADH dehydrogenase subunit 2 interaction and inhibiting CK2 activity, could selectively attenuate the increased NMDAR activity and neuropathic pain caused by nerve injury.

Five-year view

Using NMDAR antagonists alone may not produce adequate pain relief in the majority of patients. Moreover, NMDAR antagonists used in the clinical setting have a narrow therapeutic window, which could be improved by their combined use with other analgesic drugs. We predict that more studies will seek optimal combinations of NMDAR antagonists with other drugs such as opioids, α2-adrenoceptor agonists and anticonvulsants.

Recent understanding of the mechanisms involved in regulating NMDAR function provides new opportunities to selectively reduce the NMDAR activity associated with neuropathic pain. In the near future, more will be learned about changes in NMDAR activity in the spinal cord that are directly related to specific etiologies of neuropathic pain (e.g., traumatic nerve injury, diabetic neuropathy and postherpetic neuralgia). The only subtype-selective agents that have been developed are antagonists (e.g., ifenprodil) that selectively block NR1/NR2B receptors. Using a high-throughput screening utilizing cell lines expressing NR1/NR2A subunits and a fluorometric imaging plate reader/Ca2+ assay, several sulfonamide derivatives have been identified, which display submicromolar and micromolar potency at NR1/NR2A receptors [129]. These recent efforts to develop selective NR2A antagonists [129] could prompt more studies focusing on targeting NR2A-containing NMDARs as a treatment option for neuropathic pain. Still, the drugs targeting NR2C and NR2D subunits may also have some therapeutic potential for treatment of neuropathic pain.

In addition, several subtype-selective positive and negative allosteric modulators of NMDARs have been developed recently [130], and these new NMDAR modulators may be promising in treatments for neuropathic pain without causing severe side effects. This includes a series of naphthalene and phenanthrene derivatives that display inhibitory and/or potentiating activity with remarkably different patterns of selectivity at NMDARs containing different NR2 subunits. In brief, several compounds have been shown to selectively potentiate responses at NR1/NR2A (UBP512), NR1/NR2A and NR1/NR2B (UBP710), NR1/NR2D (UBP551) or NR1/NR2C and NR1/NR2D receptors (NSC339614) and have no effect or inhibit responses at the other NMDARs [130]. Furthermore, UBP512 only inhibits NR1/NR2C and NR1/NR2D receptors, whereas UBP608 inhibits NR1/NR2A receptors with a 23-fold selectivity compared with NR1/NR2D receptors [130]. These agents represent a novel class of NMDAR allosteric modulator drugs that do not act at the glutamate- or glycine-binding sites, the ion channel, or the N-terminal regulatory domain, but appear to act at the dimer interface between individual subunit ligand-binding domains. Therefore, these new compounds could become valuable tools in identifying the physiological roles of distinct NMDAR subtypes.

Key issues.

  • Increased N-methyl-d-aspartate receptor (NMDAR) activity at the spinal level plays an important role in the development of neuropathic pain caused by peripheral nerve injury. However, increased NMDAR activity has not been demonstrated specifically in other types of neuropathic pain, such as diabetic neuropathy and postherpetic neuralgia.

  • Increased NMDAR activity in the dorsal horn of the spinal cord after a nerve injury probably results from increased phosphorylation of NMDARs. However, the possibility of a NMDAR subunit switch from NR2B to NR2A should be investigated further in neuropathic pain conditions.

  • Disrupting specific NMDAR phosphorylation sites or inhibiting Src and CK2 may inhibit pathological NMDAR activity without blocking the physiological function of NMDARs.

  • Patients with neuropathic pain are not a homogeneous population, and treatments should be tailored to the specific etiologies and underlying mechanisms of their neuropathic pain. For example, NMDAR antagonists such as ketamine and dextromethorphan would be used to treat complex regional pain syndrome but not postherpetic neuralgia.

  • Because of the complex mechanisms involved in neuropathic pain, optimal combinations of NMDAR antagonists with other availableanalgesics could improve the relief of neuropathic pain and minimize adverse effects.

Footnotes

For reprint orders, please contact reprints@expert-reviews.com

Financial & competing interests disclosure

The authors are supported by grants (GM064830, NS045602 and NS073935) from the National Institutes of Health and the NG and Helen T Hawkins Endowment (to Hui-Lin Pan). The authors have no other relevant affiliations or financial involvement with any organization or entity with a financial interest in or financial conflict with the subject matter or materials discussed in the manuscript apart from those disclosed.

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

References

Papers of special note have been highlighted as:

• of interest

•• of considerable interest

  • 1.American Pain Foundation. Overview of American pain surveys: 2005–2006. J Pain Palliat Care Pharmacother. 2008;22:33–38. doi: 10.1080/15360280801989344. [DOI] [PubMed] [Google Scholar]
  • 2.Hollmann M, Boulter J, Maron C, et al. Zinc potentiates agonist-induced currents at certain splice variants of the NMDA receptor. Neuron. 1993;10:943–954. doi: 10.1016/0896-6273(93)90209-a. [DOI] [PubMed] [Google Scholar]
  • 3.Kutsuwada T, Kashiwabuchi N, Mori H, et al. Molecular diversity of the NMDA receptor channel. Nature. 1992;358:36–41. doi: 10.1038/358036a0. [DOI] [PubMed] [Google Scholar]
  • 4••.Monyer H, Sprengel R, Schoepfer R, et al. Heteromeric NMDA receptors: molecular and functional distinction of subtypes. Science. 1992;256:1217–1221. doi: 10.1126/science.256.5060.1217. Describes different NR2 subunits and their functions. [DOI] [PubMed] [Google Scholar]
  • 5.Ciabarra AM, Sullivan JM, Gahn LG, et al. Cloning and characterization of chi-1: a developmentally regulated member of a novel class of the ionotropic glutamate receptor family. J Neurosci. 1995;15:6498–6508. doi: 10.1523/JNEUROSCI.15-10-06498.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Sun L, Margolis FL, Shipley MT, Lidow MS. Identification of a long variant of mRNA encoding the NR3 subunit of the NMDA receptor: its regional distribution and developmental expression in the rat brain. FEBS Lett. 1998;441:392–396. doi: 10.1016/s0014-5793(98)01590-7. [DOI] [PubMed] [Google Scholar]
  • 7.Nishi M, Hinds H, Lu HP, Kawata M, Hayashi Y. Motoneuron-specific expression of NR3B, a novel NMDA-type glutamate receptor subunit that works in a dominant-negative manner. J Neurosci. 2001;21:RC185. doi: 10.1523/JNEUROSCI.21-23-j0003.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Chatterton JE, Awobuluyi M, Premkumar LS, et al. Excitatory glycine receptors containing the NR3 family of NMDA receptor subunits. Nature. 2002;415:793–798. doi: 10.1038/nature715. [DOI] [PubMed] [Google Scholar]
  • 9••.Vicini S, Wang JF, Li JH, et al. Functional and pharmacological differences between recombinant N-methyl-D-aspartate receptors. J Neurophysiol. 1998;79:555–566. doi: 10.1152/jn.1998.79.2.555. Describes distinct functional properties of NR2 subunits. [DOI] [PubMed] [Google Scholar]
  • 10.Monyer H, Burnashev N, Laurie DJ, Sakmann B, Seeburg PH. Developmental and regional expression in the rat brain and functional properties of four NMDA receptors. Neuron. 1994;12:529–540. doi: 10.1016/0896-6273(94)90210-0. [DOI] [PubMed] [Google Scholar]
  • 11.Sheng M, Cummings J, Roldan LA, Jan YN, Jan LY. Changing subunit composition of heteromeric NMDA receptors during development of rat cortex. Nature. 1994;368:144–147. doi: 10.1038/368144a0. [DOI] [PubMed] [Google Scholar]
  • 12.Laurie DJ, Bartke I, Schoepfer R, Naujoks K, Seeburg PH. Regional, developmental and interspecies expression of the four NMDAR2 subunits, examined using monoclonal antibodies. Brain Res Mol Brain Res. 1997;51:23–32. doi: 10.1016/s0169-328x(97)00206-4. [DOI] [PubMed] [Google Scholar]
  • 13.Stocca G, Vicini S. Increased contribution of NR2A subunit to synaptic NMDA receptors in developing rat cortical neurons. J Physiol. 1998;507:13–24. doi: 10.1111/j.1469-7793.1998.013bu.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Tovar KR, Westbrook GL. The incorporation of NMDA receptors with a distinct subunit composition at nascent hippocampal synapses in vitro. J Neurosci. 1999;19:4180–4188. doi: 10.1523/JNEUROSCI.19-10-04180.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Townsend M, Yoshii A, Mishina M, Constantine-Paton M. Developmental loss of miniature N-methyl-D-aspartate receptor currents in NR2A knockout mice. Proc Natl Acad Sci USA. 2003;100:1340–1345. doi: 10.1073/pnas.0335786100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Momiyama A. Distinct synaptic and extrasynaptic NMDA receptors identified in dorsal horn neurones of the adult rat spinal cord. J Physiol. 2000;523:621–628. doi: 10.1111/j.1469-7793.2000.t01-1-00621.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Thompson CL, Drewery DL, Atkins HD, Stephenson FA, Chazot PL. Immunohistochemical localization of N-methyl-D-aspartate receptor subunits in the adult murine hippocampal formation: evidence for a unique role of the NR2D subunit. Brain Res Mol Brain Res. 2002;102:55–61. doi: 10.1016/s0169-328x(02)00183-3. [DOI] [PubMed] [Google Scholar]
  • 18.Yung KK. Localization of glutamate receptors in dorsal horn of rat spinal cord. Neuroreport. 1998;9:1639–1644. doi: 10.1097/00001756-199805110-00069. [DOI] [PubMed] [Google Scholar]
  • 19.Boyce S, Wyatt A, Webb JK, et al. Selective NMDA NR2B antagonists induce antinociception without motor dysfunction: correlation with restricted localisation of NR2B subunit in dorsal horn. Neuropharmacology. 1999;38:611–623. doi: 10.1016/s0028-3908(98)00218-4. [DOI] [PubMed] [Google Scholar]
  • 20.Nagy GG, Watanabe M, Fukaya M, Todd AJ. Synaptic distribution of the NR1, NR2A and NR2B subunits of the N-methyl-D-aspartate receptor in the rat lumbar spinal cord revealed with an antigen-unmasking technique. Eur J Neurosci. 2004;20:3301–3312. doi: 10.1111/j.1460-9568.2004.03798.x. [DOI] [PubMed] [Google Scholar]
  • 21.Sato K, Kiyama H, Park HT, Tohyama M. AMPA, KA and NMDA receptors are expressed in the rat DRG neurones. Neuroreport. 1993;4:1263–1265. doi: 10.1097/00001756-199309000-00013. [DOI] [PubMed] [Google Scholar]
  • 22.Liu H, Wang H, Sheng M, et al. Evidence for presynaptic N-methyl-D-aspartate autoreceptors in the spinal cord dorsal horn. Proc Natl Acad Sci USA. 1994;91:8383–8387. doi: 10.1073/pnas.91.18.8383. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Marvizon JC, McRoberts JA, Ennes HS, et al. Two N-methyl-D-aspartate receptors in rat dorsal root ganglia with different subunit composition and localization. J Comp Neurol. 2002;446:325–341. doi: 10.1002/cne.10202. [DOI] [PubMed] [Google Scholar]
  • 24.Li J, McRoberts JA, Nie J, Ennes HS, Mayer EA. Electrophysiological characterization of N-methyl-D-aspartate receptors in rat dorsal root ganglia neurons. Pain. 2004;109:443–452. doi: 10.1016/j.pain.2004.02.021. [DOI] [PubMed] [Google Scholar]
  • 25.Matsuda K, Kamiya Y, Matsuda S, Yuzaki M. Cloning and characterization of a novel NMDA receptor subunit NR3B: a dominant subunit that reduces calcium permeability. Brain Res Mol Brain Res. 2002;100:43–52. doi: 10.1016/s0169-328x(02)00173-0. [DOI] [PubMed] [Google Scholar]
  • 26.Perez-Otano I, Schulteis CT, Contractor A, et al. Assembly with the NR1 subunit is required for surface expression of NR3A-containing NMDA receptors. J Neurosci. 2001;21:1228–1237. doi: 10.1523/JNEUROSCI.21-04-01228.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Matsuda K, Fletcher M, Kamiya Y, Yuzaki M. Specific assembly with the NMDA receptor 3B subunit controls surface expression and calcium permeability of NMDA receptors. J Neurosci. 2003;23:10064–10073. doi: 10.1523/JNEUROSCI.23-31-10064.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Wong HK, Liu XB, Matos MF, et al. Temporal and regional expression of NMDA receptor subunit NR3A in the mammalian brain. J Comp Neurol. 2002;450:303–317. doi: 10.1002/cne.10314. [DOI] [PubMed] [Google Scholar]
  • 29.Prithviraj R, Inglis FM. Expression of the N-methyl-D-aspartate receptor subunit NR3B regulates dendrite morphogenesis in spinal motor neurons. Neuroscience. 2008;155:145–153. doi: 10.1016/j.neuroscience.2008.03.089. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Wee KS, Zhang Y, Khanna S, Low CM. Immunolocalization of NMDA receptor subunit NR3B in selected structures in the rat forebrain, cerebellum, and lumbar spinal cord. J Comp Neurol. 2008;509:118–135. doi: 10.1002/cne.21747. [DOI] [PubMed] [Google Scholar]
  • 31.Mayer ML, Westbrook GL, Guthrie PB. Voltage-dependent block by Mg2+ of NMDA responses in spinal cord neurones. Nature. 1984;309:261–263. doi: 10.1038/309261a0. [DOI] [PubMed] [Google Scholar]
  • 32.Nowak L, Bregestovski P, Ascher P, Herbet A, Prochiantz A. Magnesium gates glutamate-activated channels in mouse central neurones. Nature. 1984;307:462–465. doi: 10.1038/307462a0. [DOI] [PubMed] [Google Scholar]
  • 33.Burzomato V, Frugier G, Perez-Otano I, Kittler JT, Attwell D. The receptor subunits generating NMDA receptor mediated currents in oligodendrocytes. J Physiol. 2010;588:3403–3414. doi: 10.1113/jphysiol.2010.195503. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Wollmuth LP, Kuner T, Sakmann B. Adjacent asparagines in the NR2-subunit of the NMDA receptor channel control the voltage-dependent block by extracellular Mg2+ J Physiol. 1998;506:13–32. doi: 10.1111/j.1469-7793.1998.013bx.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Bardoni R, Magherini PC, MacDermott AB. NMDA EPSCs at glutamatergic synapses in the spinal cord dorsal horn of the postnatal rat. J Neurosci. 1998;18:6558–6567. doi: 10.1523/JNEUROSCI.18-16-06558.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Zhou HY, Zhang HM, Chen SR, Pan HL. Increased C-fiber nociceptive input potentiates inhibitory glycinergic transmission in the spinal dorsal horn. J Pharmacol Exp Ther. 2008;324:1000–1010. doi: 10.1124/jpet.107.133470. [DOI] [PubMed] [Google Scholar]
  • 37.Raman IM, Tong G, Jahr CE. β-adrenergic regulation of synaptic NMDA receptors by cAMP-dependent protein kinase. Neuron. 1996;16:415–421. doi: 10.1016/s0896-6273(00)80059-8. [DOI] [PubMed] [Google Scholar]
  • 38.Tingley WG, Ehlers MD, Kameyama K, et al. Characterization of protein kinase A and protein kinase C phosphorylation of the N-methyl-D-aspartate receptor NR1 subunit using phosphorylation site-specific antibodies. J Biol Chem. 1997;272:5157–5166. doi: 10.1074/jbc.272.8.5157. [DOI] [PubMed] [Google Scholar]
  • 39.Leonard AS, Hell JW. Cyclic AMP-dependent protein kinase and protein kinase C phosphorylate N-methyl-D-aspartate receptors at different sites. J Biol Chem. 1997;272:12107–12115. doi: 10.1074/jbc.272.18.12107. [DOI] [PubMed] [Google Scholar]
  • 40.Chen L, Huang LY. Protein kinase C reduces Mg2+ block of NMDA-receptor channels as a mechanism of modulation. Nature. 1992;356:521–523. doi: 10.1038/356521a0. [DOI] [PubMed] [Google Scholar]
  • 41.Lieberman DN, Mody I. Casein kinase-II regulates NMDA channel function in hippocampal neurons. Nat Neurosci. 1999;2:125–132. doi: 10.1038/5680. [DOI] [PubMed] [Google Scholar]
  • 42.Chung HJ, Huang YH, Lau LF, Huganir RL. Regulation of the NMDA receptor complex and trafficking by activity-dependent phosphorylation of the NR2B subunit PDZ ligand. J Neurosci. 2004;24:10248–10259. doi: 10.1523/JNEUROSCI.0546-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Leonard AS, Lim IA, Hemsworth DE, Horne MC, Hell JW. Calcium/calmodulin-dependent protein kinase II is associated with the N-methyl-D-aspartate receptor. Proc Natl Acad Sci USA. 1999;96:3239–3244. doi: 10.1073/pnas.96.6.3239. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Barria A, Malinow R. NMDA receptor subunit composition controls synaptic plasticity by regulating binding to CaMKII. Neuron. 2005;48:289–301. doi: 10.1016/j.neuron.2005.08.034. [DOI] [PubMed] [Google Scholar]
  • 45.Yu XM, Askalan R, Keil GJ, 2nd, Salter MW. NMDA channel regulation by channel-associated protein tyrosine kinase Src. Science. 1997;275:674–678. doi: 10.1126/science.275.5300.674. [DOI] [PubMed] [Google Scholar]
  • 46.Kohr G, Seeburg PH. Subtype-specific regulation of recombinant NMDA receptor-channels by protein tyrosine kinases of the Src family. J Physiol. 1996;492:445–452. doi: 10.1113/jphysiol.1996.sp021320. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Wechsler A, Teichberg VI. Brain spectrin binding to the NMDA receptor is regulated by phosphorylation, calcium and calmodulin. EMBO J. 1998;17:3931–3939. doi: 10.1093/emboj/17.14.3931. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Abe T, Matsumura S, Katano T, et al. Fyn kinase-mediated phosphorylation of NMDA receptor NR2B subunit at Tyr1472 is essential for maintenance of neuropathic pain. Eur J Neurosci. 2005;22:1445–1454. doi: 10.1111/j.1460-9568.2005.04340.x. [DOI] [PubMed] [Google Scholar]
  • 49.Lu WY, Xiong ZG, Lei S, et al. G-protein-coupled receptors act via protein kinase C and Src to regulate NMDA receptors. Nat Neurosci. 1999;2:331–338. doi: 10.1038/7243. [DOI] [PubMed] [Google Scholar]
  • 50.Markram H, Segal M. Activation of protein kinase C suppresses responses to NMDA in rat CA1 hippocampal neurones. J Physiol. 1992;457:491–501. doi: 10.1113/jphysiol.1992.sp019389. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Raymond LA, Tingley WG, Blackstone CD, Roche KW, Huganir RL. Glutamate receptor modulation by protein phosphorylation. J Physiol (Paris) 1994;88:181–192. doi: 10.1016/0928-4257(94)90004-3. [DOI] [PubMed] [Google Scholar]
  • 52.Huang Y, Lu W, Ali DW, et al. CAKβ/Pyk2 kinase is a signaling link for induction of long-term potentiation in CA1 hippocampus. Neuron. 2001;29:485–496. doi: 10.1016/s0896-6273(01)00220-3. [DOI] [PubMed] [Google Scholar]
  • 53.Jones ML, Leonard JP. PKC site mutations reveal differential modulation by insulin of NMDA receptors containing NR2A or NR2B subunits. J Neurochem. 2005;92:1431–1438. doi: 10.1111/j.1471-4159.2004.02985.x. [DOI] [PubMed] [Google Scholar]
  • 54.Liao GY, Wagner DA, Hsu MH, Leonard JP. Evidence for direct protein kinase-C mediated modulation of N-methyl-D-aspartate receptor current. Mol Pharmacol. 2001;59:960–964. doi: 10.1124/mol.59.5.960. [DOI] [PubMed] [Google Scholar]
  • 55.Crump FT, Dillman KS, Craig AM. cAMP-dependent protein kinase mediates activity-regulated synaptic targeting of NMDA receptors. J Neurosci. 2001;21:5079–5088. doi: 10.1523/JNEUROSCI.21-14-05079.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Omkumar RV, Kiely MJ, Rosenstein AJ, Min KT, Kennedy MB. Identification of a phosphorylation site for calcium/calmodulin dependent protein kinase II in the NR2B subunit of the N-methyl-D-aspartate receptor. J Biol Chem. 1996;271:31670–31678. doi: 10.1074/jbc.271.49.31670. [DOI] [PubMed] [Google Scholar]
  • 57.Wang YT, Salter MW. Regulation of NMDA receptors by tyrosine kinases and phosphatases. Nature. 1994;369:233–235. doi: 10.1038/369233a0. [DOI] [PubMed] [Google Scholar]
  • 58.Lu YM, Roder JC, Davidow J, Salter MW. Src activation in the induction of long-term potentiation in CA1 hippocampal neurons. Science. 1998;279:1363–1367. doi: 10.1126/science.279.5355.1363. [DOI] [PubMed] [Google Scholar]
  • 59.Zheng F, Gingrich MB, Traynelis SF, Conn PJ. Tyrosine kinase potentiates NMDA receptor currents by reducing tonic zinc inhibition. Nat Neurosci. 1998;1:185–191. doi: 10.1038/634. [DOI] [PubMed] [Google Scholar]
  • 60.Lau LF, Huganir RL. Differential tyrosine phosphorylation of N-methyl-D-aspartate receptor subunits. J Biol Chem. 1995;270:20036–20041. doi: 10.1074/jbc.270.34.20036. [DOI] [PubMed] [Google Scholar]
  • 61.Yang M, Leonard JP. Identification of mouse NMDA receptor subunit NR2A C-terminal tyrosine sites phosphorylated by coexpression with v-Src. J Neurochem. 2001;77:580–588. doi: 10.1046/j.1471-4159.2001.00255.x. [DOI] [PubMed] [Google Scholar]
  • 62.Moon IS, Apperson ML, Kennedy MB. The major tyrosine-phosphorylated protein in the postsynaptic density fraction is N-methyl-D-aspartate receptor subunit 2B. Proc Natl Acad Sci USA. 1994;91:3954–3958. doi: 10.1073/pnas.91.9.3954. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Rostas JA, Brent VA, Voss K, et al. Enhanced tyrosine phosphorylation of the 2B subunit of the N-methyl-D-aspartate receptor in long-term potentiation. Proc Natl Acad Sci USA. 1996;93:10452–10456. doi: 10.1073/pnas.93.19.10452. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Lin SY, Wu K, Levine ES, et al. BDNF acutely increases tyrosine phosphorylation of the NMDA receptor subunit 2B in cortical and hippocampal postsynaptic densities. Brain Res Mol Brain Res. 1998;55:20–27. doi: 10.1016/s0169-328x(97)00349-5. [DOI] [PubMed] [Google Scholar]
  • 65.Nakazawa T, Komai S, Tezuka T, et al. Characterization of Fyn-mediated tyrosine phosphorylation sites on GluR ε 2 (NR2B) subunit of the N-methyl-D-aspartate receptor. J Biol Chem. 2001;276:693–699. doi: 10.1074/jbc.M008085200. [DOI] [PubMed] [Google Scholar]
  • 66.Chen C, Leonard JP. Protein tyrosine kinase-mediated potentiation of currents from cloned NMDA receptors. J Neurochem. 1996;67:194–200. doi: 10.1046/j.1471-4159.1996.67010194.x. [DOI] [PubMed] [Google Scholar]
  • 67.Charriaut-Marlangue C, Otani S, Creuzet C, Ben-Ari Y, Loeb J. Rapid activation of hippocampal casein kinase II during long-term potentiation. Proc Natl Acad Sci USA. 1991;88:10232–10236. doi: 10.1073/pnas.88.22.10232. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Kimura R, Matsuki N. Protein kinase CK2 modulates synaptic plasticity by modification of synaptic NMDA receptors in the hippocampus. J Physiol. 2008;586:3195–3206. doi: 10.1113/jphysiol.2008.151894. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Soto D, Pancetti F, Marengo JJ, et al. Protein kinase CK2 in postsynaptic densities: phosphorylation of PSD-95/SAP90 and NMDA receptor regulation. Biochem Biophys Res Commun. 2004;322:542–550. doi: 10.1016/j.bbrc.2004.07.158. [DOI] [PubMed] [Google Scholar]
  • 70••.Sanz-Clemente A, Matta JA, Isaac JT, Roche KW. Casein kinase 2 regulates the NR2 subunit composition of synaptic NMDA receptors. Neuron. 2010;67:984–996. doi: 10.1016/j.neuron.2010.08.011. Demonstrates CK2 regulation of NR2A/NR2B trafficking. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Wang LY, Orser BA, Brautigan DL, MacDonald JF. Regulation of NMDA receptors in cultured hippocampal neurons by protein phosphatases 1 and 2A. Nature. 1994;369:230–232. doi: 10.1038/369230a0. [DOI] [PubMed] [Google Scholar]
  • 72.Lieberman DN, Mody I. Regulation of NMDA channel function by endogenous Ca2+-dependent phosphatase. Nature. 1994;369:235–239. doi: 10.1038/369235a0. [DOI] [PubMed] [Google Scholar]
  • 73.Wang YT, Yu XM, Salter MW. Ca2+-independent reduction of N-methyl-D-aspartate channel activity by protein tyrosine phosphatase. Proc Natl Acad Sci USA. 1996;93:1721–1725. doi: 10.1073/pnas.93.4.1721. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Zhang HM, Chen SR, Pan HL. Effects of activation of group III metabotropic glutamate receptors on spinal synaptic transmission in a rat model of neuropathic pain. Neuroscience. 2009;158:875–884. doi: 10.1016/j.neuroscience.2008.10.042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Li JQ, Chen SR, Chen H, Cai YQ, Pan HL. Regulation of increased glutamatergic input to spinal dorsal horn neurons by mGluR5 in diabetic neuropathic pain. J Neurochem. 2010;112:162–172. doi: 10.1111/j.1471-4159.2009.06437.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Wang XL, Zhang HM, Chen SR, Pan HL. Altered synaptic input and GABAB receptor function in spinal superficial dorsal horn neurons in rats with diabetic neuropathy. J Physiol. 2007;579:849–861. doi: 10.1113/jphysiol.2006.126102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Zhou HY, Chen SR, Chen H, Pan HL. Functional plasticity of group II metabotropic glutamate receptors in regulating spinal excitatory and inhibitory synaptic input in neuropathic pain. J Pharmacol Exp Ther. 2011;336:254–264. doi: 10.1124/jpet.110.173112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78•.Liu H, Mantyh PW, Basbaum AI. NMDA-receptor regulation of substance P release from primary afferent nociceptors. Nature. 1997;386:721–724. doi: 10.1038/386721a0. Demonstrates the presence of presynaptic N-methyl-D-aspartate receptors in the spinal cord dorsal horn. [DOI] [PubMed] [Google Scholar]
  • 79.Zhou HY, Chen SR, Chen H, Pan HL. Opioid-induced long-term potentiation in the spinal cord is a presynaptic event. J Neurosci. 2010;30:4460–4466. doi: 10.1523/JNEUROSCI.5857-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Bardoni R, Torsney C, Tong CK, Prandini M, MacDermott AB. Presynaptic NMDA receptors modulate glutamate release from primary sensory neurons in rat spinal cord dorsal horn. J Neurosci. 2004;24:2774–2781. doi: 10.1523/JNEUROSCI.4637-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Wu ZZ, Chen SR, Pan HL. Transient receptor potential vanilloid type 1 activation down-regulates voltage-gated calcium channels through calcium-dependent calcineurin in sensory neurons. J Biol Chem. 2005;280:18142–18151. doi: 10.1074/jbc.M501229200. [DOI] [PubMed] [Google Scholar]
  • 82•.Isaev D, Gerber G, Park SK, Chung JM, Randik M. Facilitation of NMDA-induced currents and Ca2+ transients in the rat substantia gelatinosa neurons after ligation of L5–L6 spinal nerves. Neuroreport. 2000;11:4055–4061. doi: 10.1097/00001756-200012180-00030. Demonstrates increased N-methyl-D-aspartate currents of spinal dorsal horn neurons by nerve injury. [DOI] [PubMed] [Google Scholar]
  • 83.Zhou HY, Zhang HM, Chen SR, Pan HL. Protein kinase CK2 increases synaptic NMDA receptor activity in the spinal cord in neuropathic pain. Neurosci Abstr. 2009;39(613.3):B88. [Google Scholar]
  • 84.Ultenius C, Linderoth B, Meyerson BA, Wallin J. Spinal NMDA receptor phosphorylation correlates with the presence of neuropathic signs following peripheral nerve injury in the rat. Neurosci Lett. 2006;399:85–90. doi: 10.1016/j.neulet.2006.01.018. [DOI] [PubMed] [Google Scholar]
  • 85•.Gao X, Kim HK, Chung JM, Chung K. Enhancement of NMDA receptor phosphorylation of the spinal dorsal horn and nucleus gracilis neurons in neuropathic rats. Pain. 2005;116:62–72. doi: 10.1016/j.pain.2005.03.045. Demonstrates increased phosphorylation of the NR1 subunit in the spinal dorsal horn after nerve injury. [DOI] [PubMed] [Google Scholar]
  • 86•.Liu XJ, Gingrich JR, Vargas-Caballero M, et al. Treatment of inflammatory and neuropathic pain by uncoupling Src from the NMDA receptor complex. Nat Med. 2008;14:1325–1332. doi: 10.1038/nm.1883. Demonstrates the reversal of neuropathic pain by disrupting the Src NADH dehydrogenase subunit 2 interaction in an animal model. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Herrero JF, Laird JM, Lopez-Garcia JA. Wind-up of spinal cord neurones and pain sensation: much ado about something? Prog Neurobiol. 2000;61:169–203. doi: 10.1016/s0301-0082(99)00051-9. [DOI] [PubMed] [Google Scholar]
  • 88.Davies SN, Lodge D. Evidence for involvement of N-methylaspartate receptors in ‘wind-up’ of class 2 neurones in the dorsal horn of the rat. Brain Res. 1987;424:402–406. doi: 10.1016/0006-8993(87)91487-9. [DOI] [PubMed] [Google Scholar]
  • 89.Dickenson AH, Sullivan AF. Evidence for a role of the NMDA receptor in the frequency dependent potentiation of deep rat dorsal horn nociceptive neurones following C fibre stimulation. Neuropharmacology. 1987;26:1235–1238. doi: 10.1016/0028-3908(87)90275-9. [DOI] [PubMed] [Google Scholar]
  • 90.Suzuki R, Matthews EA, Dickenson AH. Comparison of the effects of MK-801, ketamine and memantine on responses of spinal dorsal horn neurones in a rat model of mononeuropathy. Pain. 2001;91:101–109. doi: 10.1016/s0304-3959(00)00423-1. [DOI] [PubMed] [Google Scholar]
  • 91.Iwata H, Takasusuki T, Yamaguchi S, Hori Y. NMDA receptor 2B subunit-mediated synaptic transmission in the superficial dorsal horn of peripheral nerve-injured neuropathic mice. Brain Res. 2007;1135:92–101. doi: 10.1016/j.brainres.2006.12.014. [DOI] [PubMed] [Google Scholar]
  • 92.Qu XX, Cai J, Li MJ, et al. Role of the spinal cord NR2B-containing NMDA receptors in the development of neuropathic pain. Exp Neurol. 2009;215:298–307. doi: 10.1016/j.expneurol.2008.10.018. [DOI] [PubMed] [Google Scholar]
  • 93.Miletic G, Pankratz MT, Miletic V. Increases in the phosphorylation of cyclic AMP response element binding protein (CREB) and decreases in the content of calcineurin accompany thermal hyperalgesia following chronic constriction injury in rats. Pain. 2002;99:493–500. doi: 10.1016/S0304-3959(02)00242-7. [DOI] [PubMed] [Google Scholar]
  • 94.Mao J, Price DD, Hayes RL, et al. Intrathecal treatment with dextrorphan or ketamine potently reduces pain-related behaviors in a rat model of peripheral mononeuropathy. Brain Res. 1993;605:164–168. doi: 10.1016/0006-8993(93)91368-3. [DOI] [PubMed] [Google Scholar]
  • 95.Eisenberg E, LaCross S, Strassman AM. The clinically tested N-methyl-D-aspartate receptor antagonist memantine blocks and reverses thermal hyperalgesia in a rat model of painful mononeuropathy. Neurosci Lett. 1995;187:17–20. doi: 10.1016/0304-3940(95)11326-r. [DOI] [PubMed] [Google Scholar]
  • 96.Mao J, Price DD, Mayer DJ, Lu J, Hayes RL. Intrathecal MK-801 and local nerve anesthesia synergistically reduce nociceptive behaviors in rats with experimental peripheral mononeuropathy. Brain Res. 1992;576:254–262. doi: 10.1016/0006-8993(92)90688-6. [DOI] [PubMed] [Google Scholar]
  • 97.Davar G, Hama A, Deykin A, Vos B, Maciewicz R. MK-801 blocks the development of thermal hyperalgesia in a rat model of experimental painful neuropathy. Brain Res. 1991;553:327–330. doi: 10.1016/0006-8993(91)90844-l. [DOI] [PubMed] [Google Scholar]
  • 98.Qian J, Brown SD, Carlton SM. Systemic ketamine attenuates nociceptive behaviors in a rat model of peripheral neuropathy. Brain Res. 1996;715:51–62. doi: 10.1016/0006-8993(95)01452-7. [DOI] [PubMed] [Google Scholar]
  • 99.Chaplan SR, Malmberg AB, Yaksh TL. Efficacy of spinal NMDA receptor antagonism in formalin hyperalgesia and nerve injury evoked allodynia in the rat. J Pharmacol Exp Ther. 1997;280:829–838. [PubMed] [Google Scholar]
  • 100.Bennett AD, Everhart AW, Hulsebosch CE. Intrathecal administration of an NMDA or a non-NMDA receptor antagonist reduces mechanical but not thermal allodynia in a rodent model of chronic central pain after spinal cord injury. Brain Res. 2000;859:72–82. doi: 10.1016/s0006-8993(99)02483-x. [DOI] [PubMed] [Google Scholar]
  • 101.Holtman JR, Jr, Crooks PA, Johnson-Hardy JK, et al. Effects of norketamine enantiomers in rodent models of persistent pain. Pharmacol Biochem Behav. 2008;90:676–685. doi: 10.1016/j.pbb.2008.05.011. [DOI] [PubMed] [Google Scholar]
  • 102•.Chen SR, Samoriski G, Pan HL. Antinociceptive effects of chronic administration of uncompetitive NMDA receptor antagonists in a rat model of diabetic neuropathic pain. Neuropharmacology. 2009;57:121–126. doi: 10.1016/j.neuropharm.2009.04.010. Demonstrates the analgesic effect of chronic administration of memantine or neramexane in an animal model of painful diabetic neuropathy. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Christoph T, Reissmuller E, Schiene K, Englberger W, Chizh BA. Antiallodynic effects of NMDA glycine(B) antagonists in neuropathic pain: possible peripheral mechanisms. Brain Res. 2005;1048:218–227. doi: 10.1016/j.brainres.2005.04.081. [DOI] [PubMed] [Google Scholar]
  • 104.Christoph T, Schiene K, Englberger W, Parsons CG, Chizh BA. The antiallodynic effect of NMDA antagonists in neuropathic pain outlasts the duration of the in vivo NMDA antagonism. Neuropharmacology. 2006;51:12–17. doi: 10.1016/j.neuropharm.2006.02.007. [DOI] [PubMed] [Google Scholar]
  • 105.Quartaroli M, Carignani C, Dal Forno G, et al. Potent antihyperalgesic activity without tolerance produced by glycine site antagonist of N-methyl-D-aspartate receptor GV196771A. J Pharmacol Exp Ther. 1999;290:158–169. [PubMed] [Google Scholar]
  • 106.Quartaroli M, Fasdelli N, Bettelini L, Maraia G, Corsi M. GV196771A, an NMDA receptor/glycine site antagonist, attenuates mechanical allodynia in neuropathic rats and reduces tolerance induced by morphine in mice. Eur J Pharmacol. 2001;430:219–227. doi: 10.1016/s0014-2999(01)01278-x. [DOI] [PubMed] [Google Scholar]
  • 107.Bare TM, Brown DG, Horchler CL, et al. Pyridazinoquinolinetriones as NMDA glycine-site antagonists with oral antinociceptive activity in a model of neuropathic pain. J Med Chem. 2007;50:3113–3131. doi: 10.1021/jm060212s. [DOI] [PubMed] [Google Scholar]
  • 108.Zhang W, Shi CX, Gu XP, Ma ZL, Zhu W. Ifenprodil induced antinociception and decreased the expression of NR2B subunits in the dorsal horn after chronic dorsal root ganglia compression in rats. Anesth Analg. 2009;108:1015–1020. doi: 10.1213/ane.0b013e318193ffd2. [DOI] [PubMed] [Google Scholar]
  • 109.Gu X, Zhang J, Ma Z, et al. The role of N-methyl-D-aspartate receptor subunit NR2B in spinal cord in cancer pain. Eur J Pain. 2010;14:496–502. doi: 10.1016/j.ejpain.2009.09.001. [DOI] [PubMed] [Google Scholar]
  • 110.Nakazato E, Kato A, Watanabe S. Brain but not spinal NR2B receptor is responsible for the anti-allodynic effect of an NR2B subunit-selective antagonist CP-101, 606 in a rat chronic constriction injury model. Pharmacology. 2005;73:8–14. doi: 10.1159/000081069. [DOI] [PubMed] [Google Scholar]
  • 111.Gingrich JR, Pelkey KA, Fam SR, et al. Unique domain anchoring of Src to synaptic NMDA receptors via the mitochondrial protein NADH dehydrogenase subunit 2. Proc Natl Acad Sci USA. 2004;101:6237–6242. doi: 10.1073/pnas.0401413101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Correll GE, Maleki J, Gracely EJ, Muir JJ, Harbut RE. Subanesthetic ketamine infusion therapy: a retrospective analysis of a novel therapeutic approach to complex regional pain syndrome. Pain Med. 2004;5:263–275. doi: 10.1111/j.1526-4637.2004.04043.x. [DOI] [PubMed] [Google Scholar]
  • 113••.Kiefer RT, Rohr P, Ploppa A, et al. Efficacy of ketamine in anesthetic dosage for the treatment of refractory complex regional pain syndrome: an open-label Phase II study. Pain Med. 2008;9:1173–1201. doi: 10.1111/j.1526-4637.2007.00402.x. Demonstrates significant pain relief by ketamine infusion in patients with complex regional pain syndrome. [DOI] [PubMed] [Google Scholar]
  • 114.Duale C, Sibaud F, Guastella V, et al. Perioperative ketamine does not prevent chronic pain after thoracotomy. Eur J Pain. 2009;13:497–505. doi: 10.1016/j.ejpain.2008.06.013. [DOI] [PubMed] [Google Scholar]
  • 115.Huge V, Lauchart M, Magerl W, et al. Effects of low-dose intranasal (S)-ketamine in patients with neuropathic pain. Eur J Pain. 2010;14:387–394. doi: 10.1016/j.ejpain.2009.08.002. [DOI] [PubMed] [Google Scholar]
  • 116••.Nelson KA, Park KM, Robinovitz E, Tsigos C, Max MB. High-dose oral dextromethorphan versus placebo in painful diabetic neuropathy and postherpetic neuralgia. Neurology. 1997;48:1212–1218. doi: 10.1212/wnl.48.5.1212. Demonstrates that dextromethorphan is effective in the treatment of painful diabetic neuropathy, but not postherpetic neuralgia, in patients. [DOI] [PubMed] [Google Scholar]
  • 117.Pud D, Eisenberg E, Spitzer A, et al. The NMDA receptor antagonist amantadine reduces surgical neuropathic pain in cancer patients: a double blind, randomized, placebo controlled trial. Pain. 1998;75:349–354. doi: 10.1016/s0304-3959(98)00014-1. [DOI] [PubMed] [Google Scholar]
  • 118.Fukui S, Komoda Y, Nosaka S. Clinical application of amantadine, an NMDA antagonist, for neuropathic pain. J Anesth. 2001;15:179–181. doi: 10.1007/s005400170025. [DOI] [PubMed] [Google Scholar]
  • 119••.Schwartzman RJ, Alexander GM, Grothusen JR, et al. Outpatient intravenous ketamine for the treatment of complex regional pain syndrome: a double-blind placebo controlled study. Pain. 2009;147:107–115. doi: 10.1016/j.pain.2009.08.015. Demonstrates the reduction of pain by ketamine infusion in patients with complex regional pain syndrome. [DOI] [PubMed] [Google Scholar]
  • 120••.Sigtermans MJ, van Hilten JJ, Bauer MC, et al. Ketamine produces effective and long-term pain relief in patients with complex regional pain syndrome type 1. Pain. 2009;145:304–311. doi: 10.1016/j.pain.2009.06.023. Demonstrates the long-term pain relief after ketamine treatment in patients with complex regional pain syndrome. [DOI] [PubMed] [Google Scholar]
  • 121.Leung A, Wallace MS, Ridgeway B, Yaksh T. Concentration–effect relationship of intravenous alfentanil and ketamine on peripheral neurosensory thresholds, allodynia and hyperalgesia of neuropathic pain. Pain. 2001;91:177–187. doi: 10.1016/s0304-3959(00)00433-4. [DOI] [PubMed] [Google Scholar]
  • 122.Sinis N, Birbaumer N, Gustin S, et al. Memantine treatment of complex regional pain syndrome: a preliminary report of six cases. Clin J Pain. 2007;23:237–243. doi: 10.1097/AJP.0b013e31802f67a7. [DOI] [PubMed] [Google Scholar]
  • 123••.Sang CN, Booher S, Gilron I, Parada S, Max MB. Dextromethorphan and memantine in painful diabetic neuropathy and postherpetic neuralgia: efficacy and dose-response trials. Anesthesiology. 2002;96:1053–1061. doi: 10.1097/00000542-200205000-00005. Shows that dextromethorphan is effective in the treatment of painful diabetic neuropathy, but not postherpetic neuralgia, in patients. [DOI] [PubMed] [Google Scholar]
  • 124.Wiech K, Kiefer RT, Topfner S, et al. A placebo-controlled randomized crossover trial of the N-methyl-D-aspartic acid receptor antagonist, memantine, in patients with chronic phantom limb pain. Anesth Analg. 2004;98:408–413. doi: 10.1213/01.ANE.0000096002.53818.BD. [DOI] [PubMed] [Google Scholar]
  • 125.Schifitto G, Yiannoutsos CT, Simpson DM, et al. A placebo-controlled study of memantine for the treatment of human immunodeficiency virus-associated sensory neuropathy. J Neurovirol. 2006;12:328–331. doi: 10.1080/13550280600873835. [DOI] [PubMed] [Google Scholar]
  • 126.Wallace MS, Rowbotham MC, Katz NP, et al. A randomized, double-blind, placebo-controlled trial of a glycine antagonist in neuropathic pain. Neurology. 2002;59:1694–1700. doi: 10.1212/01.wnl.0000036273.98213.34. [DOI] [PubMed] [Google Scholar]
  • 127.Amr YM. Multi-day low dose ketamine infusion as adjuvant to oral gabapentin in spinal cord injury related chronic pain: a prospective, randomized, double blind trial. Pain Physician. 2010;13:245–249. [PubMed] [Google Scholar]
  • 128.Mercadante S, Arcuri E, Tirelli W, Casuccio A. Analgesic effect of intravenous ketamine in cancer patients on morphine therapy: a randomized, controlled, double-blind, crossover, double-dose study. J Pain Symptom Manage. 2000;20:246–252. doi: 10.1016/s0885-3924(00)00194-9. [DOI] [PubMed] [Google Scholar]
  • 129.Bettini E, Sava A, Griffante C, et al. Identification and characterisation of novel NMDA receptor antagonists selective for NR2A- over NR2B-containing receptors. J Pharmacol Exp Ther. 2010;335:636–644. doi: 10.1124/jpet.110.172544. [DOI] [PubMed] [Google Scholar]
  • 130.Costa BM, Irvine MW, Feng G, et al. A novel family of negative and positive allosteric modulators of NMDA receptors. J Pharmacol Exp Ther. 2010;335:614–621. doi: 10.1124/jpet.110.174144. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES