Skip to main content
PMC Canada Author Manuscripts logoLink to PMC Canada Author Manuscripts
. Author manuscript; available in PMC: 2013 Aug 20.
Published in final edited form as: Neuron Glia Biol. 2012 May 22;7(1):99–108. doi: 10.1017/S1740925X12000087

Brain-Derived Neurotrophic Factor from Microglia: A Molecular Substrate for Neuropathic Pain

Tuan Trang 1, Simon Beggs 1, Michael W Salter 1
PMCID: PMC3748035  CAMSID: CAMS3159  PMID: 22613083

Abstract

One of the most significant advances in pain research is the realization that neurons are not the only cell type involved in the etiology of chronic pain. This realization has caused a radical shift from the previous dogma that neuronal dysfunction alone accounts for pain pathologies, to the current framework of thinking that takes into account all cell types within the central nervous system (CNS). This shift in thinking stems from growing evidence that glia can modulate the function and directly shape the cellular architecture of nociceptive networks in the CNS. Microglia, in particular, are increasingly recognized as active principal players that respond to changes in physiological homeostasis by extending their processes toward the site of neural damage, and by releasing specific factors that have profound consequences on neuronal function and that contribute to CNS pathologies caused by disease or injury. A key molecule that modulates microglia activity is ATP, an endogenous ligand of the P2 receptor family. Microglia express several P2 receptor subtypes, and of these the P2X4 receptor subtype has emerged as a core microglia-neuron signaling pathway: activation of this receptor drives the release of brain-derived neurotrophic factor (BDNF), a cellular substrate that causes disinhibition of pain-transmitting spinal lamina I neurons. Converging evidence points to BDNF from spinal microglia as being a critical microglia-neuron signalling molecule that gates aberrant nociceptive processing in the spinal cord. The present review highlights recent advances in our understanding of P2X4 receptor-mediated signaling and regulation of BDNF in microglia, as well as the implications for microglia-neuron interactions in the pathobiology of neuropathic pain.

Keywords: Microglia, Brain-derived neurotrophic factor, Purinoceptors, Neuropathic pain, Nerve injury, Chronic pain

INTRODUCTION

Acute nociceptive pain is instructive and warns against imminent or existing tissue damage, whereas chronic pain has no known defensive or beneficial function and remains one of the most vexing challenges in medicine. In the United States alone, the direct economic costs of acute and chronic pain conditions are estimated to cost $560–635 billion annually (Institute of Medicine Report, 2011). Although major advances have been made towards understanding the fundamental processes that underlie acute pain, little is known about the cellular and molecular processes initiated by acute pain that lead to the progression to chronic pain, which is characterized by pain persisting long after the tissue damage has healed. Paradoxically, the degree of injury is not necessarily predictive of the severity or chronicity of the pain, and the development of chronic pain following injury is highly variable between individuals. Considerable evidence suggests that transition from acute to chronic pain is a consequence of alterations in the cellular, molecular, and anatomical organization of nociceptive neural networks in the spinal dorsal horn and brain (Latremoliere and Woolf, 2009;Scholz and Woolf, 2002;Voscopoulos and Lema, 2010;Woolf and Salter, 2000). Thus, an acute injury can have profound effects on the nociceptive neural circuitry, transforming the CNS from conveying normally protective acute nociceptive signals into a pathologically altered system in which pain can occur spontaneously and responses to innocuous and noxious stimuli are amplified.

Among the most severe and debilitating types of chronic pain conditions is neuropathic pain which can arise following lesions to the somatosensory nervous system caused by trauma, infection, or pathology (Scholz and Woolf, 2002;Zimmermann, 2001;Gwak and Hulsebosch, 2009). Such damage to a nerve can instigate a series of cellular and molecular changes that directly affect neuronal plasticity, leading ultimately to altered synaptic connectivity and the reorganization of peripheral and central nociceptive circuitry (Scholz and Woolf, 2002;Woolf and Salter, 2000;Latremoliere and Woolf, 2009). These changes have profound effects on pain transmission by suppressing mechanisms that inhibit pain and by enhancing mechanisms that facilitate pain in the central nervous system (CNS). The consequent shift in inhibitory and excitatory control can cause a pathological amplification that alters the modality of sensory input and output from the spinal cord to elicit exaggerated pain responses that typifies chronic pain conditions (Costigan et al., 2009).

A principal locus of this pathologically altered activity arises from the nociceptive neurons in lamina I of the spinal dorsal horn. The action potential discharge of these neurons is normally evoked only in response to noxious peripheral stimulation (Keller et al., 2007). However, after peripheral nerve injury the output of lamina I neurons is transformed such that innocuous stimulus can evoke action potential discharges, and response to noxious stimulation is greatly exaggerated. In addition, after peripheral nerve injury lamina I neurons exhibit spontaneous bursting in the absence of overt stimulation whereas in uninjured animals these neurons are normally silent. These changes in the output of lamina I neurons provide a neural basis for the three cardinal signs of neuropathic pain in humans (Woolf and Salter, 2000): mechanical allodynia (discharge in response to innocuous stimulation), hyperalgesia (exaggerated response to noxious stimulation), and spontaneous pain (spontaneous bursting in the absence of an overt stimulus). The symptoms of neuropathic pain are often resistant to the current available treatments, which have typically been directed against cellular targets in neurons. Failure of the current cadre of drugs in treating the sequelae of neuropathic pain has intensified the search for new molecular players and cellular substrates that can be targeted for development of novel pharmacological therapies. This search has generated a rapidly growing body of evidence that indicates interactions between neurons and glia are critical in establishing and maintaining neuropathic pain (Beggs and Salter, 2010;Grace et al., 2011;Inoue and Tsuda, 2006;Milligan and Watkins, 2009;Gwak and Hulsebosch, 2010). Microglia in particular, have emerged as key players in the initiation and in the expression of neuropathic pain (Inoue and Tsuda, 2006;Trang et al., 2011;Tsuda et al., 2003;Tsuda et al., 2005;Watkins et al., 2001;Watkins and Maier, 2003).

Spinal microglia respond to peripheral nerve injury

Microglia originate from embryonic macrophages derived from the yolk sac during late prenatal development (Ginhoux et al., 2010). In the adult CNS, microglia comprise 5–10% of the total glial population (Kreutzberg, 1996;Lawson et al., 1990;Nakajima and Kohsaka, 2001). Likened to the electricians of the CNS, microglia are able to modulate the electrical activity within neuronal circuits (Graeber, 2010). In the surveillance mode of activity, microglia possess small soma bearing thin ramified processes that cover large non-overlapping territories throughout the brain and spinal cord (Bushong et al., 2002;Kreutzberg, 1996). In response to stimuli that potentially disrupt homeostasis, microglia rapidly extend their processes toward the site of neural damage, forming a barrier between healthy and injured cells that limits damage in the CNS (Davalos et al., 2005;Nimmerjahn et al., 2005). The rapid microglia responsiveness precedes a more slowly developing series of changes in morphology, gene expression, function, and proliferation (Hanisch and Kettenmann, 2007;Kettenmann et al., 2011;Kreutzberg, 1996;Nakajima and Kohsaka, 2001). These changes are observed in spinal microglia in rodent models of peripheral nerve injury caused by compression, ligation, or transaction (Calvo and Bennett, 2011;Calvo et al., 2011;Echeverry et al., 2008;Liu et al., 1995;Tsuda et al., 2003;Zhang and De Koninck, 2006). The stereotypical microglial response also entails the upregulation of surface marker proteins belonging to the complement cascade: complement receptor 3 (CR3), Toll-like receptor 4 (TLR4), CD14, CD4, and major histocompatibility complex (MHC) class I and II (Coyle, 1998;Liu et al., 1995;Sweitzer et al., 2002;Tanga et al., 2004;Tsuda et al., 2003). Upregulation of these surface proteins concomitant with distinct anatomical changes in morphology – from ramified to ameboid – are stereotypical of microglia in the ‘enhanced response state’. Although microglia in this state have been reported to release of a myriad of factors that signal to neurons in the spinal cord to alter neuronal excitability, in pain neuroplasticity after peripheral nerve injury a specific core pathway has been implicated through P2X4 receptor-stimulated release of BDNF (Beggs and Salter, 2010;Trang et al., 2011).

Spinal microglia express P2 receptors

A critical molecular substrate for microglia-neuron communication is ATP, an endogenous ligand of the P2 receptor family comprising of P2Y metabotropic and P2X ionotropic receptors (Coull et al., 2005;Di, 2006;Jarvis, 2010;Maeda et al., 2010). Microglia express a variety of P2 receptors: of the metabotropic P2Y receptors, microglia express P2Y1, 2, 4, 6, and 12 receptors (Boucsein et al., 2003;Farber and Kettenmann, 2005;Inoue and Tsuda, 2006;Sasaki et al., 2003), whereas microglial expression of the ionotropic P2X receptor is restricted to the P2X4 and P2X7R subtypes (Inoue and Tsuda, 2006;Collo et al., 1997;Ferrari et al., 1996;Tsuda et al., 2003). Despite the diverse repertoire of P2Y receptors expressed on microglia, only P2Y12 receptors have been demonstrably implicated in the development of tactile allodynia associated with peripheral nerve injury (Kobayashi et al., 2008;Tozaki-Saitoh et al., 2008). By contrast, both P2X4 and P2X7 receptors are causally involved in expressing nerve injury-induced pain behaviours (Chessell et al., 2005;Coull et al., 2005;Kobayashi et al., 2008;Tozaki-Saitoh et al., 2008;Tsuda et al., 2003;Tsuda et al., 2009a;Tsuda et al., 2009c;Ulmann et al., 2008).

A compelling argument has been made for involvement of P2X7 receptors in inflammatory pain and in neuropathic pain on the basis of reduced pain sensitivity in P2X7 receptor-deficient mice (Chessell et al., 2005) and after pharmacological blockade of the receptor (Broom et al., 2008;Dell’Antonio et al., 2002;Honore et al., 2006;Honore et al., 2009;McGaraughty et al., 2007;Perez-Medrano et al., 2009). Recent evidence has also established a linked between genetic variations in the P2X7 receptor that affect its function and variability in chronic pain sensitivity in mice and in humans (Sorge et al., In Press). The specific cell type(s) responsible for the P2X7 receptor mediated pain phenotypes has been difficult to precisely pinpoint because, in addition to being expressed on microglia, P2X7 receptors are localized on macrophages, neurons, and astrocytes (Donnelly-Roberts et al., 2008;Donnelly-Roberts and Jarvis, 2007;Skaper et al., 2010). Thus, conclusions about the specific role microglial P2X7Rs play in inflammatory and neuropathic pain is confounded by expression of this receptor on a wide variety of cell types. By contrast, much of our conceptual understanding of the role microglial P2 receptors play in neuropathic pain stems from elucidating the fundamental mechanisms that regulate P2X4 receptor expression, as well as the identification of signaling pathways downstream from this receptor.

Microglial P2X4 receptors are critically required for neuropathic pain

The first compelling clues that identified P2X4 receptors as a critical molecular component in neuropathic pain were the discoveries that blocking P2X4 receptor function pharmacologically or suppressing its expression with antisense RNA transiently reverses mechanical allodynia (Tsuda et al., 2003). The development of mechanical allodynia was found to correlate temporally with an increase in spinal P2X4 receptor expression, and unexpectedly this increase was confined to microglia in the ipsilateral spinal dorsal horn. These observations were subsequently confirmed in CX3CR1+/GFP mice, in which induction of P2X4 receptors resulting from peripheral nerve lesion is restricted to activated eGFP expressing spinal microglia (Ulmann et al., 2008), and in mice lacking the P2X4 receptor, which do not develop mechanical allodynia after peripheral nerve injury (Tsuda et al., 2009a;Ulmann et al., 2008). Although neuropathic pain behaviours in the P2X4 receptor deficient mice are absent, the microglial proliferative response and the alterations in microglia morphology induced by peripheral nerve injury were not affected (Tsuda et al., 2003; Ulmann et al., 2008), suggesting that while tonic P2X4 receptor activation is required for maintaining peripheral nerve injury-induced allodynia, the proliferation and upregulation of microglial P2X4 receptors in the spinal cord are mediated by distinct intracellular mechanisms. Direct evidence that stimulation of P2X4 receptors expressed on microglia is sufficient to elicit pain hypersensitivity comes from the finding that injection of P2X4 receptor-stimulated cultured microglia into the spinal cords of naïve animals elicits robust mechanical allodynia that is blocked by 2′,3′-O-(2,4,6-trinitrophenyl)adenosine 5′-triphosphate (TNP-ATP) (Coull et al., 2005;Tsuda et al., 2003;Tsuda et al., 2008b). Taken together, the pharmacological, genetic, and behavioral findings indicate that activity of P2X4 receptors expressed on spinal microglia is critically involved in the functional alterations in the spinal dorsal horn that maintain ongoing pain following peripheral nerve injury.

Regulation of P2X4 receptor expression in microglia

A major question arising from the observation that development of mechanical hypersensitivity is correlated with a progressive increase in spinal P2X4 receptor expression is how peripheral nerve injury initiates signalling in the spinal dorsal horn to specifically cause an increase in P2X4 receptor expression in microglia. The answer to this question appears to involve the release of several signalling elements including: CCL21, a chemokine released from injured neurons that functions as an upstream activator of P2X4 receptor (Biber et al., 2011;de Jong et al., 2005), interferon γ, a cytokine that transforms resting spinal microglia into an activated state (Tsuda et al., 2009b), and tryptase, a protease released from mast cells that activates proteinase-activated receptor 2 in microglia (Yuan et al., 2010). Also critical for upregulating expression of P2X4 receptors is the extracellular matrix molecule fibronectin, which through activity of Lyn kinase and downstream activation of intracellular signalling pathways involving phosphatidylinositol 3-kinase (PI3K)-Akt and mitogen-activated protein kinase kinase (MAPK kinase, MEK)-extracellular signal-regulated kinase (ERK), modulates the transcriptional and post-transcriptional levels of P2X4 receptor expression in microglia (Nasu-Tada et al., 2006;Tsuda et al., 2008a;Tsuda et al., 2008b;Tsuda et al., 2009c). Thus, several elements of the molecular machinery required for upregulation of P2X4 receptors in microglia following peripheral nerve injury have recently been identified (Figure 1). The implications of this diverse modulation and whether they are causally connected through a convergent common pathway that controls P2X4 receptor expression is not known.

Figure 1.

Figure 1

Dynamic regulation of P2X4 receptors in microglia. Microglia in the physiological CNS actively monitor their surrounding environment for potential stimuli that threaten homeostasis. In response to peripheral nerve injury spinal microglia upregulate expression of P2X4 receptors, which normally are expressed at low levels in the ‘resting/surveillance state.’ Upregulation of P2X4 receptors is a critical mechanistic step through which spinal microglia signal to neurons in the spinal dorsal horn to cause neuropathic pain. Activation of P2X4 receptors initiates the p38 MAPK-BDNF-KCC2 signalling cascade to cause aberrant nociceptive output that underlies pain hypersensitivity characterized by hyperalgesia, allodynia, and spontaneous pain. Molecules released from injured neurons, such as the chemokines CCL2 and CCL21, as well as the cytokine IFN-γ, increase P2X4 receptor expression in microglia. The fibronectin-Lyn kinase signalling cascade and tryptase released from mast cells have also been found to upregulate P2X4 receptors in microglia. Thus, the P2X4 receptor and the components of its signalling pathway have emerged as a core mechanism through which spinal microglia contribute to neuropathic pain.

Significant inroads have also been made in understanding the intracellular trafficking of P2X4 receptors. It is now known that the proportion of P2X4 receptors on the cell surface is regulated by rapid constitutive internalizationand reinsertion into the plasma membrane (Bobanovic et al., 2002;Fujii et al., 2011;Royle et al., 2002;Toulme et al., 2006). Internalization is controlled by the C-terminus of the P2X4 receptor (Fujii et al., 2011;Qureshi et al., 2007;Royle et al., 2002), a region also important for agonist-induced desensitization (Fountain and North, 2006) and phosphoinositide PIP2 modulation of P2X4 receptor function (Bernier et al., 2008). In microglia, internalized P2X4 receptors are targeted to lysosomes which contain a large pool of P2X4 receptors. Mobilization of P2X4 receptors contained within these lysosomes to, as well as the retrieval of these receptors from, the plasma membrane regulates the proportion of P2X4 receptors expressed on the cell surface (Qureshi et al., 2007). Dynamic targeting of P2X4 receptor containing lysosomes in microglia can be induced by the chemokine CCL2, acting via the CCR2 chemokine receptor (Toyomitsu et al., 2012). Moreover, a rise in intracellular Ca2+ is sufficient to trigger lysosome exocytosis that leads to accumulation of P2X4 receptors on the cell surface and enhancement of P2X4 receptor-mediated currents (Qureshi et al., 2007).

Activation of cell surface P2X4 receptors by ATP causes structural changes that result in two distinct conformations – a prototypical cation channel or a macropore structure. In the presence of extracellular Ca2+, ATP stimulation induces the P2X4 receptor to transiently open a non-selective cation permeable channel, but in the absence of extracellular Ca2+ the receptor forms a macropore that allows passage of large molecules (Shinozaki et al., 2009). P2X4 receptors expressed in microglia possess this ability to function both as a cation channel and as a macropore (Bernier et al., 2010;Seil et al., 2010); however, the implication of this dual mode of functioning in the context of neuropathic pain remains an intriguingly open question. The discovery of two distinct P2X4 receptor modes of function, in addition to identification of the P2X4 receptor crystal structure, have illuminated many atomic details about its extracellular domain structure, the putative ATP binding site, transmembrane regions, and ion permeation pathway (Kawate et al., 2009).

P2X4R activation drives release of brain-derived neurotrophic factor

Another key question arising from the discovery that microglial P2X4 receptors are critically involved in neuropathic pain is – how does P2X4R signaling in microglia affect nociceptive processing in the spinal dorsal horn? It was reasoned that P2X4 receptors in microglia must initiate signalling that is communicated to neurons in the spinal nociceptive network which then relay the information to the brain (Tsuda et al., 2003;Tsuda et al., 2005). It was predicted that such signaling could occur through the release of one or more diffusible chemical messengers from microglia upon stimulation of P2X4 receptors. The components of the microglia-neuron signalling pathway was elucidated by the discovery that activation of P2X4 receptors in microglia evokes the release of brain-derived neurotrophic factor (BDNF) which fundamentally alters the output of spinal lamina I neurons to the brain (Coull et al., 2005;De Koninck, 2007;Keller et al., 2007). BDNF was found to down-regulate expression of the K+-Cl co-transporter KCC2, the main Cl transporter in spinal lamina I neurons, causing a rise in intracellular [Cl] within these cells (Coull et al., 2003;Coull et al., 2005). With the increase in [Cl], opening of GABAA or glycine channels was less effective in producing inhibition, and in approximately one-third of lamina I neurons GABA-evoked responses were converted from hyperpolarizing to depolarizing (Coull et al., 2005).

Several key lines of evidence support BDNF as being the critical microglia-neuron signalling molecule: 1) disrupting BDNF-TrkB signalling with TrkB antibody, or sequestering BDNF with TrkB-Fc fusion protein, prevents mechanical allodynia evoked by administering P2X4 receptor-stimulated microglia; 2) siRNA knockdown of BDNF prevents the effects of intrathecally administered microglia on lamina I neurons and on the microglia-elicited pain behaviours; and 3) ATP activation of P2X4 receptors causes BDNF release from microglia in culture, a response prevented by TNP-ATP or by knocking down expression of BDNF with siRNA (Coull et al., 2005). Taken together, the most parsimonious explanation for these findings is that P2X4 receptor stimulated microglia signal to spinal lamina I nociceptive neurons causing aberrant spinal nociceptive processing, and that the critical microglia-neuron signaling molecule is BDNF. The requirement for P2X4 receptors in the release of BDNF is consistent with observations that P2X4 receptor-deficient mice have impaired microglial BDNF release, possess altered BDNF signaling in the spinal cord, and they are protected from developing mechanical allodynia following peripheral nerve injury (Ulmann et al., 2008). Indeed, there is overwhelming evidence for BDNF involvement in the initiation of central sensitization associated with neuropathic pain (Biggs et al., 2010;Lever et al., 2003;Lu et al., 2007;Obata et al., 2011). However, the previous conclusions that BDNF derived from primary afferent neurons is entirely responsible for spinal nociceptive hypersensitivity has been brought into question by evidence indicating that there is a lack of primary afferent evoked BDNF release in the spinal cord after nerve injury (Lever et al., 2003), and that eliminating BDNF from primary afferents suppresses inflammatory pain but has no effect on nerve injury-induced mechanical allodynia (Zhao et al., 2006). Together, these findings raise the possibility that neuronal-derived and microglial-derived BDNF have distinct roles – BDNF from neurons may be required for inflammatory pain, whereas BDNF from microglia may mediate neuropathic pain.

Microglial P2X4 receptors signal through Ca2+ and p38-MAPK

ATP stimulation of P2X4 receptors instigates a series of conformational changes that allow cations, such as Ca2+ and Na+, entry into the cell through a non-selective channel (Burnstock, 2006a;Burnstock, 2006b;North, 2002). In primary microglia culture, influx of extracellular Ca2+ is necessary for P2X4 receptor-stimulated release of pre-existing and newly synthesized BDNF (Trang et al., 2009;Ulmann et al., 2008). Ca2+-dependency is a hallmark of release by exocytosis, and in microglia, P2X4 receptor-stimulated release of BDNF involves a Ca2+- and SNARE (soluble N-ethylmaleimide-sensitive factor attachment protein receptor)-dependent vesicular exocytotic pathway (Trang et al., 2009). This release mechanism is consistent with evidence that SNARE proteins SNAP-23, syntaxin-1, and cellubrevin are expressed in microglial cells (Hepp et al., 1999).

Ca2+-dependent release of BDNF is well-characterized in neurons and occurs from both pre-synaptic terminals and post-synaptic dendrites (Balkowiec and Katz, 2002;Kolarow et al., 2007). In nerve terminals, the release of BDNF is triggered by depolarization that depends upon Ca2+ influx through voltage-gated Ca2+ channels, and upon release of Ca2+ from intracellular stores (Buldyrev et al., 2006;Lever et al., 2003). Release of BDNF from post-synaptic dendrites is also mediated by influx of Ca2+, initiated by opening NMDA receptors or voltage-gated Ca2+ channels (Kolarow et al., 2007). By contrast, BDNF release from microglia appears to be independent of intracellular Ca2+ stores (Trang et al., 2009).

In addition to causing release of BDNF, P2X4 receptor-stimulated Ca2+ influx drives the transcription and translation of BDNF which results in accumulation of the neurotrophin in the microglia (Trang et al., 2009). BDNF was originally identified in neurons as an immediate-early gene (Lauterborn et al., 1996;West et al., 2001), transcription of which is initiated by Ca2+-influx through L-type voltage-gated Ca2+ channels or via NMDA receptors (Ghosh et al., 1994;Kolarow et al., 2007;Tao et al., 1998). Thus, Ca2+-dependent stimulation of BDNF transcription is common to neurons and microglia. In neurons, stimulus-evoked increase in Ca2+ activates key transcription factors that regulate the BDNF gene, such as cAMP response element binding protein (Tao et al., 1998), nuclear factor kappaB (Lipsky et al., 2001;Marini et al., 2004), methyl-CpG-binding protein 2 (Chen et al., 2003;Zhou et al., 2006), and calcium-responsive transcription factor (Tao et al., 1998). Whether the P2X4 receptor-stimulated increase in BDNF transcription in microglia involves these transcription factors, or others, is as yet undetermined.

In neurons, expression of BDNF is mediated by the extracellular signal-regulated protein kinase (ERK) and p38-MAPK signaling pathways (Ji and Woolf, 2001;Obata et al., 2004;Rao et al., 2007). Activation of these pathways in spinal microglia contributes to the development of neuropathic pain behaviors in nerve-injured rats (Jin et al., 2003;Tsuda et al., 2004). However, in primary microglia cultures isolated from the rat brain, pharmacological inhibitors of ERK have no effect on the release or the accumulation of BDNF, indicating ERK is not required in the P2X4 receptor-BDNF signaling pathway in microglia. By contrast, inhibiting activity of p38-MAPK prevents both the P2X4 receptor-evoked accumulation and release of BDNF (Trang et al., 2009). This finding suggests that activation of p38-MAPK by ATP is necessary for the BDNF response, and together with the Ca2+-dependence of p38-MAPK activation, the simplest explanation is that influx of Ca2+ through the P2X4 receptor is a critical step linking stimulation of these receptors to the p38-MAPK signalling pathway. By demonstrating that p38-MAPK is a cellular intermediary in the BDNF response, a unifying mechanism in microglia can account for ongoing expression of neuropathic pain behaviors as requiring activity of both P2X4 receptor and p38-MAPK.

Ca2+-dependent activation of p38-MAPK has been reported in a number of cell types (Blanquet, 2000;Dehez et al., 2001), including peripheral macrophages (Ulmann et al., 2010). Like microglia, macrophages are immune cells that constitutively express functional P2X4 receptors (Brone et al., 2007;Qureshi et al., 2007;Ulmann et al., 2010). A recent study has uncovered that stimulating P2X4 receptors expressed in macrophages triggers Ca2+ influx and p38-MAPK activation leading to the production and release of prostaglandin E2 (Ulmann et al., 2010), a principal substrate for peripheral inflammation (Portanova et al., 1996;Samad et al., 2002). p38-MAPK signalling therefore appears to gate the release of distinct signaling molecules from microglia (BDNF) and peripheral macrophages (prostaglandin E2) that have distinct roles in the pathoetiology of neuropathic pain and inflammatory pain, respectively.

Convergence of P2X7 and P2Y12 receptor activation on p38-MAPK signalling

In addition to P2X4 receptors, P2X7 and P2Y12 receptors expressed on microglia are increasingly implicated in neuropathic pain (Clark et al., 2007;Clark et al., 2010;Kobayashi et al., 2008;Tozaki-Saitoh et al., 2008). Activation of P2X7 receptors can cause release of interleukin-1β and cathepsin S, molecules which in the spinal cord may contribute to mechanical hypersensitivity following injury to a peripheral nerve (Clark et al., 2007;Clark et al., 2010). Analogous to P2X4 receptor-mediated release of BDNF, the release of interleukin-1β and cathepsin S from microglia requires P2X7 receptor signaling to p38-MAPK (Clark et al., 2010). Likewise, involvement of microglial P2Y12 receptors in neuropathic pain critically depends on p38-MAPK activation (Kobayashi et al., 2008); however, the signaling events downstream from P2Y12 receptor activation of p38-MAPK have yet to be elucidated. Collectively, these findings suggest that p38-MAPK is a cellular intermediary and its activation is a key point of convergence for P2X4, P2X7, and P2Y12 receptor signaling in neuropathic pain. The next major challenges are to determine the significance of this convergence in signalling for microglia function, and how the interplay of upstream and downstream components in these signalling pathways contributes to neuropathic pain.

Conclusions

Conventional neurocentric bias has focused almost entirely on neuron-to-neuron signalling as being the fundamental basis of neuropathic pain. However, this view fails to taken into account the diversity of cell types and the intricacies of cell-cell interactions in the CNS. In this respect, there has been a seismic shift towards understanding pathologies in the CNS as being a consequence of the complex interplay between diverse cell types involving multiple convergent and divergent signalling pathways. There is now a canon of literature detailing a plethora of molecular links between neurons and glia and their involvement in the pathogenesis of pain hypersensitivity arising from peripheral nerve injury. Microglia, in particular, have emerged as key cellular targets for treating neuropathic pain (Beggs and Salter, 2010;Calvo and Bennett, 2011;Scholz and Woolf, 2002;Trang et al., 2006). In the search for drugs directed against microglia, caution must be heeded in developing ‘microglia inhibitors’ with non-specific glial modulating properties given that microglia may play both a protective and a pathological role (Rivest et al., 2009; Milligan and Watkins, 2009). Rather than attempting to globally inhibit microglia, an approach might be to specifically target key microglia-neuron signalling pathways. As described above, the P2X4 receptor is a cellular hub through which microglia transform the nociceptive output of the spinal dorsal horn. Therefore, strategies directed against the P2X4 receptor signalling pathway is a high value target for the potential treatment of neuropathic pain caused by peripheral nerve injury. Moreover, elucidation of the core P2X4 receptor signalling pathway opens the possibility that this pathway, or its components, may be a common mechanism underlying other pathological pain states in addition to neuropathic pain.

Major advances have been made in identifying the essential molecular components that regulate P2X4 receptor expression and the cellular machinery that drives release of microglial BDNF, a critical substrate for neuropathic pain (Figure 2). These discoveries have built a complete framework for understanding microglia-to-neuron signaling in neuropathic pain, positioning the P2X4 receptor as a core signaling pathway necessary for ongoing expression of tactile allodynia following nerve injury.

Figure 2.

Figure 2

ATP activation of P2X4 receptors causes the synthesis and release of BDNF from microglia through a Ca2+ and p38-MAPK dependent signalling cascade. Activation of P2X4 receptors expressed on the cell surface instigates a series of conformational changes that allow cations, such as Ca2+ and Na+, entry into the cell through a non-selective channel. In a recent study, we reported ATP stimulation causes a biphasic release of BDNF and a concomitant increase in BDNF protein level within primary cultured microglia (Trang et al., 2009). Although drug treatment affected both BDNF accumulation and release, points 1–4 only illustrate the effects on the peak BDNF release response at 60 min post ATP stimulation: 1) In the presence of extracellular Ca2+, activating P2X4 receptors with ATP causes the release of BDNF; 2) Treatment with the p38-MAPK inhibitor SB203580 prevents BDNF release; 3) Inhibiting transcription or translation with actinomycin-D or cycloheximide, respectively, prevents BDNF release; and 4) Microglia treated with TAT-NSF, which interferes with NSF hexamerization necessary for SNARE-dependent exocytosis, abolishes BDNF release. The current model is that influx of Ca2+ through P2X4 receptors drives the activation of p38-MAPK that leads to increase in the SNARE-dependent release per se of BDNF and to an increase in BDNF synthesis in microglia. Modified from Trang et al., 2009.

A key cellular intermediary in the P2X4 receptor signalling pathway is p38-MAPK. Activation of this kinase is a point of convergence for involvement of microglial P2X7 and P2Y12 receptors in neuropathic pain, and for macrophage P2X4 receptor-mediate inflammatory pain. However, in contrast to the P2X4 receptor signalling pathway which has been systematically dissected, the upstream and downstream components of the P2X7 and P2Y12 receptor pathways have yet to be uncovered. Thus, the next major challenges are to determine the significance of this convergence in signaling for microglia function and how the components of P2X7 and P2Y12 receptor signalling contribute to aberrant spinal nociceptive processing in neuropathic pain.

Acknowledgments

The work of the authors is supported by grants from the Canadian Institutes of Health Research (CIHR; grant number MT-11219), the Krembil Foundation, and the Ontario Research Fund Research Excellence Program. MWS holds a Canada Research Chair (Tier I) in Neuroplasticity and Pain, and is the Anne and Max Tanenbaum Chair in Molecular Medicine at the Hospital for Sick Children. TT was supported by a CIHR fellowship.

Footnotes

Statement of Interest: None

Reference List

  1. Balkowiec A, Katz DM. Cellular mechanisms regulating activity-dependent release of native brain-derived neurotrophic factor from hippocampal neurons. J Neurosci. 2002;22:10399–10407. doi: 10.1523/JNEUROSCI.22-23-10399.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Beggs S, Salter MW. Microglia-neuronal signalling in neuropathic pain hypersensitivity 2.0. Curr Opin Neurobiol. 2010;20:474–480. doi: 10.1016/j.conb.2010.08.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Bernier LP, Boue-Grabot E, Seguela P. Functional modulation of P2X4 receptor-channels by UDP-activated P2Y6 receptors. 2010. [Google Scholar]
  4. Bernier LP, Ase AR, Chevallier S, Blais D, Zhao Q, Boue-Grabot E, Logothetis D, Seguela P. Phosphoinositides regulate P2X4 ATP-gated channels through direct interactions. J Neurosci. 2008;28:12938–12945. doi: 10.1523/JNEUROSCI.3038-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Biber K, Tsuda M, Tozaki-Saitoh H, Tsukamoto K, Toyomitsu E, Masuda T, Boddeke H, Inoue K. Neuronal CCL21 up-regulates microglia P2X4 expression and initiates neuropathic pain development. EMBO J. 2011 doi: 10.1038/emboj.2011.89. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Biggs JE, Lu VB, Stebbing MJ, Balasubramanyan S, Smith PA. Is BDNF sufficient for information transfer between microglia and dorsal horn neurons during the onset of central sensitization? Mol Pain. 2010;6:44. doi: 10.1186/1744-8069-6-44. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Blanquet PR. Identification of two persistently activated neurotrophin-regulated pathways in rat hippocampus. Neuroscience. 2000;95:705–719. doi: 10.1016/s0306-4522(99)00489-3. [DOI] [PubMed] [Google Scholar]
  8. Bobanovic LK, Royle SJ, Murrell-Lagnado RD. P2X receptor trafficking in neurons is subunit specific. J Neurosci. 2002;22:4814–4824. doi: 10.1523/JNEUROSCI.22-12-04814.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Boucsein C, Zacharias R, Farber K, Pavlovic S, Hanisch UK, Kettenmann H. Purinergic receptors on microglial cells: functional expression in acute brain slices and modulation of microglial activation in vitro. Eur J Neurosci. 2003;17:2267–2276. doi: 10.1046/j.1460-9568.2003.02663.x. [DOI] [PubMed] [Google Scholar]
  10. Brone B, Moechars D, Marrannes R, Mercken M, Meert T. P2X currents in peritoneal macrophages of wild type and P2X4 −/− mice. Immunol Lett. 2007;113:83–89. doi: 10.1016/j.imlet.2007.07.015. [DOI] [PubMed] [Google Scholar]
  11. Broom DC, Matson DJ, Bradshaw E, Buck ME, Meade R, Coombs S, Matchett M, Ford KK, Yu W, Yuan J, Sun SH, Ochoa R, Krause JE, Wustrow DJ, Cortright DN. Characterization of N-(adamantan-1-ylmethyl)-5-[(3R-amino-pyrrolidin-1-yl)methyl]-2-chloro-ben zamide, a P2X7 antagonist in animal models of pain and inflammation. J Pharmacol Exp Ther. 2008;327:620–633. doi: 10.1124/jpet.108.141853. [DOI] [PubMed] [Google Scholar]
  12. Buldyrev I, Tanner NM, Hsieh HY, Dodd EG, Nguyen LT, Balkowiec A. Calcitonin gene-related peptide enhances release of native brain-derived neurotrophic factor from trigeminal ganglion neurons. J Neurochem. 2006;99:1338–1350. doi: 10.1111/j.1471-4159.2006.04161.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Burnstock G. Pathophysiology and therapeutic potential of purinergic signaling. Pharmacol Rev. 2006a;58:58–86. doi: 10.1124/pr.58.1.5. [DOI] [PubMed] [Google Scholar]
  14. Burnstock G. Purinergic P2 receptors as targets for novel analgesics. Pharmacol Ther. 2006b;110:433–454. doi: 10.1016/j.pharmthera.2005.08.013. [DOI] [PubMed] [Google Scholar]
  15. Bushong EA, Martone ME, Jones YZ, Ellisman MH. Protoplasmic astrocytes in CA1 stratum radiatum occupy separate anatomical domains. J Neurosci. 2002;22:183–192. doi: 10.1523/JNEUROSCI.22-01-00183.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Calvo M, Bennett DL. The mechanisms of microgliosis and pain following peripheral nerve injury. Exp Neurol. 2011 doi: 10.1016/j.expneurol.2011.08.018. [DOI] [PubMed] [Google Scholar]
  17. Calvo M, Zhu N, Grist J, Ma Z, Loeb JA, Bennett DL. Following nerve injury neuregulin-1 drives microglial proliferation and neuropathic pain via the MEK/ERK pathway. Glia. 2011;59:554–568. doi: 10.1002/glia.21124. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Chen WG, Chang Q, Lin Y, Meissner A, West AE, Griffith EC, Jaenisch R, Greenberg ME. Derepression of BDNF transcription involves calcium-dependent phosphorylation of MeCP2. Science. 2003;302:885–889. doi: 10.1126/science.1086446. [DOI] [PubMed] [Google Scholar]
  19. Chessell IP, Hatcher JP, Bountra C, Michel AD, Hughes JP, Green P, Egerton J, Murfin M, Richardson J, Peck WL, Grahames CB, Casula MA, Yiangou Y, Birch R, Anand P, Buell GN. Disruption of the P2X7 purinoceptor gene abolishes chronic inflammatory and neuropathic pain. Pain. 2005;114:386–396. doi: 10.1016/j.pain.2005.01.002. [DOI] [PubMed] [Google Scholar]
  20. Clark AK, Wodarski R, Guida F, Sasso O, Malcangio M. Cathepsin S release from primary cultured microglia is regulated by the P2X7 receptor. Glia. 2010;58:1710–1726. doi: 10.1002/glia.21042. [DOI] [PubMed] [Google Scholar]
  21. Clark AK, Yip PK, Grist J, Gentry C, Staniland AA, Marchand F, Dehvari M, Wotherspoon G, Winter J, Ullah J, Bevan S, Malcangio M. Inhibition of spinal microglial cathepsin S for the reversal of neuropathic pain. Proc Natl Acad Sci USA. 2007;104:10655–10660. doi: 10.1073/pnas.0610811104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Collo G, Neidhart S, Kawashima E, Kosco-Vilbois M, North RA, Buell G. Tissue distribution of the P2X7 receptor. Neuropharmacology. 1997;36:1277–1283. doi: 10.1016/s0028-3908(97)00140-8. [DOI] [PubMed] [Google Scholar]
  23. Costigan M, Scholz J, Woolf CJ. Neuropathic pain: a maladaptive response of the nervous system to damage. Annu Rev Neurosci. 2009;32:1–32. doi: 10.1146/annurev.neuro.051508.135531. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Coull JA, Beggs S, Boudreau D, Boivin D, Tsuda M, Inoue K, Gravel C, Salter MW, De Koninck Y. BDNF from microglia causes the shift in neuronal anion gradient underlying neuropathic pain. Nature. 2005;438:1017–1021. doi: 10.1038/nature04223. [DOI] [PubMed] [Google Scholar]
  25. Coull JA, Boudreau D, Bachand K, Prescott SA, Nault F, Sik A, De Koninck P, De Koninck Y. Trans-synaptic shift in anion gradient in spinal lamina I neurons as a mechanism of neuropathic pain. Nature. 2003;424:938–942. doi: 10.1038/nature01868. [DOI] [PubMed] [Google Scholar]
  26. Coyle DE. Partial peripheral nerve injury leads to activation of astroglia and microglia which parallels the development of allodynic behavior. Glia. 1998;23:75–83. [PubMed] [Google Scholar]
  27. Davalos D, Grutzendler J, Yang G, Kim JV, Zuo Y, Jung S, Littman DR, Dustin ML, Gan WB. ATP mediates rapid microglial response to local brain injury in vivo. Nat Neurosci. 2005;8:752–758. doi: 10.1038/nn1472. [DOI] [PubMed] [Google Scholar]
  28. de Jong EK, Dijkstra IM, Hensens M, Brouwer N, van AM, Liem RS, Boddeke HW, Biber K. Vesicle-mediated transport and release of CCL21 in endangered neurons: a possible explanation for microglia activation remote from a primary lesion. J Neurosci. 2005;25:7548–7557. doi: 10.1523/JNEUROSCI.1019-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. De Koninck Y. Altered chloride homeostasis in neurological disorders: a new target. Curr Opin Pharmacol. 2007;7:93–99. doi: 10.1016/j.coph.2006.11.005. [DOI] [PubMed] [Google Scholar]
  30. Dehez S, Daulhac L, Kowalski-Chauvel A, Fourmy D, Pradayrol L, Seva C. Gastrin-induced DNA synthesis requires p38-MAPK activation via PKC/Ca(2+) and Src-dependent mechanisms. FEBS Lett. 2001;496:25–30. doi: 10.1016/s0014-5793(01)02396-1. [DOI] [PubMed] [Google Scholar]
  31. Dell’Antonio G, Quattrini A, Cin ED, Fulgenzi A, Ferrero ME. Relief of inflammatory pain in rats by local use of the selective P2X7 ATP receptor inhibitor, oxidized ATP. Arthritis Rheum. 2002;46:3378–3385. doi: 10.1002/art.10678. [DOI] [PubMed] [Google Scholar]
  32. Di VF. Purinergic signalling between axons and microglia. Novartis Found Symp. 2006;276:253–258. doi: 10.1002/9780470032244.ch20. [DOI] [PubMed] [Google Scholar]
  33. Donnelly-Roberts D, McGaraughty S, Shieh CC, Honore P, Jarvis MF. Painful purinergic receptors. J Pharmacol Exp Ther. 2008;324:409–415. doi: 10.1124/jpet.106.105890. [DOI] [PubMed] [Google Scholar]
  34. Donnelly-Roberts DL, Jarvis MF. Discovery of P2X7 receptor-selective antagonists offers new insights into P2X7 receptor function and indicates a role in chronic pain states. Br J Pharmacol. 2007;151:571–579. doi: 10.1038/sj.bjp.0707265. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Echeverry S, Shi XQ, Zhang J. Characterization of cell proliferation in rat spinal cord following peripheral nerve injury and the relationship with neuropathic pain. Pain. 2008;135:37–47. doi: 10.1016/j.pain.2007.05.002. [DOI] [PubMed] [Google Scholar]
  36. Farber K, Kettenmann H. Physiology of microglial cells. Brain Res Brain Res Rev. 2005;48:133–143. doi: 10.1016/j.brainresrev.2004.12.003. [DOI] [PubMed] [Google Scholar]
  37. Ferrari D, Villalba M, Chiozzi P, Falzoni S, Ricciardi-Castagnoli P, Di VF. Mouse microglial cells express a plasma membrane pore gated by extracellular ATP. J Immunol. 1996;156:1531–1539. [PubMed] [Google Scholar]
  38. Fountain SJ, North RA. A C-terminal lysine that controls human P2X4 receptor desensitization. J Biol Chem. 2006;281:15044–15049. doi: 10.1074/jbc.M600442200. [DOI] [PubMed] [Google Scholar]
  39. Fujii K, Young MT, Harris KD. Exploiting powder X-ray diffraction for direct structure determination in structural biology: The P2X4 receptor trafficking motif YEQGL. J Struct Biol. 2011 doi: 10.1016/j.jsb.2011.03.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Ghosh A, Carnahan J, Greenberg ME. Requirement for BDNF in activity-dependent survival of cortical neurons. Science. 1994;263:1618–1623. doi: 10.1126/science.7907431. [DOI] [PubMed] [Google Scholar]
  41. Ginhoux F, Greter M, Leboeuf M, Nandi S, See P, Gokhan S, Mehler MF, Conway SJ, Ng LG, Stanley ER, Samokhvalov IM, Merad M. Fate mapping analysis reveals that adult microglia derive from primitive macrophages. Science. 2010;330:841–845. doi: 10.1126/science.1194637. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Grace PM, Rolan PE, Hutchinson MR. Peripheral immune contributions to the maintenance of central glial activation underlying neuropathic pain. Brain Behav Immun. 2011;25:1322–1332. doi: 10.1016/j.bbi.2011.04.003. [DOI] [PubMed] [Google Scholar]
  43. Graeber MB. Changing face of microglia. Science. 2010;330:783–788. doi: 10.1126/science.1190929. [DOI] [PubMed] [Google Scholar]
  44. Gwak YS, Hulsebosch CE. Remote astrocytic and microglial activation modulates neuronal hyperexcitability and below-level neuropathic pain after spinal injury in rat. Neuroscience. 2009;161:895–903. doi: 10.1016/j.neuroscience.2009.03.055. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Gwak YS, Hulsebosch CE. “Gliopathy” maintains persistent hyperexcitability of spinal dorsal horn neurons after spinal cord injury: substrate of central neuropathic pain. In: Costa A, Villalba E, editors. Horizons in Neuroscience Research. Vol. 1. Nova Science Publishers, Inc; 2010. pp. 195–224. [Google Scholar]
  46. Hanisch UK, Kettenmann H. Microglia: active sensor and versatile effector cells in the normal and pathologic brain. Nat Neurosci. 2007;10:1387–1394. doi: 10.1038/nn1997. [DOI] [PubMed] [Google Scholar]
  47. Hepp R, Perraut M, Chasserot-Golaz S, Galli T, Aunis D, Langley K, Grant NJ. Cultured glial cells express the SNAP-25 analogue SNAP-23. Glia. 1999;27:181–187. doi: 10.1002/(sici)1098-1136(199908)27:2<181::aid-glia8>3.0.co;2-9. [DOI] [PubMed] [Google Scholar]
  48. Honore P, Donnelly-Roberts D, Namovic M, Zhong C, Wade C, Chandran P, Zhu C, Carroll W, Perez-Medrano A, Iwakura Y, Jarvis MF. The antihyperalgesic activity of a selective P2X7 receptor antagonist, A-839977, is lost in IL-1alphabeta knockout mice. Behav Brain Res. 2009;204:77–81. doi: 10.1016/j.bbr.2009.05.018. [DOI] [PubMed] [Google Scholar]
  49. Honore P, Donnelly-Roberts D, Namovic MT, Hsieh G, Zhu CZ, Mikusa JP, Hernandez G, Zhong C, Gauvin DM, Chandran P, Harris R, Medrano AP, Carroll W, Marsh K, Sullivan JP, Faltynek CR, Jarvis MF. A-740003 [N-(1-{[(cyanoimino)(5-quinolinylamino) methyl]amino}-2,2-dimethylpropyl)-2-(3,4-dimethoxyphenyl)acetamide], a novel and selective P2X7 receptor antagonist, dose-dependently reduces neuropathic pain in the rat. J Pharmacol Exp Ther. 2006;319:1376–1385. doi: 10.1124/jpet.106.111559. [DOI] [PubMed] [Google Scholar]
  50. Inoue K, Tsuda M. The role of microglia and ATP receptors in a mechanism of neuropathic pain. Nippon Yakurigaku Zasshi. 2006;127:14–17. doi: 10.1254/fpj.127.14. [DOI] [PubMed] [Google Scholar]
  51. Jarvis MF. The neural-glial purinergic receptor ensemble in chronic pain states. Trends Neurosci. 2010;33:48–57. doi: 10.1016/j.tins.2009.10.003. [DOI] [PubMed] [Google Scholar]
  52. Ji RR, Woolf CJ. Neuronal plasticity and signal transduction in nociceptive neurons: implications for the initiation and maintenance of pathological pain. Neurobiol Dis. 2001;8:1–10. doi: 10.1006/nbdi.2000.0360. [DOI] [PubMed] [Google Scholar]
  53. Jin SX, Zhuang ZY, Woolf CJ, Ji RR. p38 mitogen-activated protein kinase is activated after a spinal nerve ligation in spinal cord microglia and dorsal root ganglion neurons and contributes to the generation of neuropathic pain. J Neurosci. 2003;23:4017–4022. doi: 10.1523/JNEUROSCI.23-10-04017.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Kawate T, Michel JC, Birdsong WT, Gouaux E. Crystal structure of the ATP-gated P2X(4) ion channel in the closed state. Nature. 2009;460:592–598. doi: 10.1038/nature08198. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Keller AF, Beggs S, Salter MW, De Koninck Y. Transformation of the output of spinal lamina I neurons after nerve injury and microglia stimulation underlying neuropathic pain. Mol Pain. 2007;3:27. doi: 10.1186/1744-8069-3-27. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Kettenmann H, Hanisch UK, Noda M, Verkhratsky A. Physiology of microglia. Physiol Rev. 2011;91:461–553. doi: 10.1152/physrev.00011.2010. [DOI] [PubMed] [Google Scholar]
  57. Kobayashi K, Yamanaka H, Fukuoka T, Dai Y, Obata K, Noguchi K. P2Y12 receptor upregulation in activated microglia is a gateway of p38 signaling and neuropathic pain. J Neurosci. 2008;28:2892–2902. doi: 10.1523/JNEUROSCI.5589-07.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Kolarow R, Brigadski T, Lessmann V. Postsynaptic secretion of BDNF and NT-3 from hippocampal neurons depends on calcium calmodulin kinase II signaling and proceeds via delayed fusion pore opening. J Neurosci. 2007;27:10350–10364. doi: 10.1523/JNEUROSCI.0692-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Kreutzberg GW. Microglia: a sensor for pathological events in the CNS. Trends Neurosci. 1996;19:312–318. doi: 10.1016/0166-2236(96)10049-7. [DOI] [PubMed] [Google Scholar]
  60. Latremoliere A, Woolf CJ. Central sensitization: a generator of pain hypersensitivity by central neural plasticity. J Pain. 2009;10:895–926. doi: 10.1016/j.jpain.2009.06.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Lauterborn JC, Rivera S, Stinis CT, Hayes VY, Isackson PJ, Gall CM. Differential effects of protein synthesis inhibition on the activity-dependent expression of BDNF transcripts: evidence for immediate-early gene responses from specific promoters. J Neurosci. 1996;16:7428–7436. doi: 10.1523/JNEUROSCI.16-23-07428.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Lawson LJ, Perry VH, Dri P, Gordon S. Heterogeneity in the distribution and morphology of microglia in the normal adult mouse brain. Neuroscience. 1990;39:151–170. doi: 10.1016/0306-4522(90)90229-w. [DOI] [PubMed] [Google Scholar]
  63. Lever I, Cunningham J, Grist J, Yip PK, Malcangio M. Release of BDNF and GABA in the dorsal horn of neuropathic rats. Eur J Neurosci. 2003;18:1169–1174. doi: 10.1046/j.1460-9568.2003.02848.x. [DOI] [PubMed] [Google Scholar]
  64. Lipsky RH, Xu K, Zhu D, Kelly C, Terhakopian A, Novelli A, Marini AM. Nuclear factor kappaB is a critical determinant in N-methyl-D-aspartate receptor-mediated neuroprotection. J Neurochem. 2001;78:254–264. doi: 10.1046/j.1471-4159.2001.00386.x. [DOI] [PubMed] [Google Scholar]
  65. Liu L, Tornqvist E, Mattsson P, Eriksson NP, Persson JK, Morgan BP, Aldskogius H, Svensson M. Complement and clusterin in the spinal cord dorsal horn and gracile nucleus following sciatic nerve injury in the adult rat. Neuroscience. 1995;68:167–179. doi: 10.1016/0306-4522(95)00103-p. [DOI] [PubMed] [Google Scholar]
  66. Lu VB, Ballanyi K, Colmers WF, Smith PA. Neuron type-specific effects of brain-derived neurotrophic factor in rat superficial dorsal horn and their relevance to ‘central sensitization’. J Physiol. 2007;584:543–563. doi: 10.1113/jphysiol.2007.141267. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Maeda M, Tsuda M, Tozaki-Saitoh H, Inoue K, Kiyama H. Nerve injury-activated microglia engulf myelinated axons in a P2Y12 signaling-dependent manner in the dorsal horn. Glia. 2010;58:1838–1846. doi: 10.1002/glia.21053. [DOI] [PubMed] [Google Scholar]
  68. Marini AM, Jiang X, Wu X, Tian F, Zhu D, Okagaki P, Lipsky RH. Role of brain-derived neurotrophic factor and NF-kappaB in neuronal plasticity and survival: From genes to phenotype. Restor Neurol Neurosci. 2004;22:121–130. [PubMed] [Google Scholar]
  69. McGaraughty S, Chu KL, Namovic MT, Donnelly-Roberts DL, Harris RR, Zhang XF, Shieh CC, Wismer CT, Zhu CZ, Gauvin DM, Fabiyi AC, Honore P, Gregg RJ, Kort ME, Nelson DW, Carroll WA, Marsh K, Faltynek CR, Jarvis MF. P2X7-related modulation of pathological nociception in rats. Neuroscience. 2007;146:1817–1828. doi: 10.1016/j.neuroscience.2007.03.035. [DOI] [PubMed] [Google Scholar]
  70. Milligan ED, Watkins LR. Pathological and protective roles of glia in chronic pain. Nat Rev Neurosci. 2009;10:23–36. doi: 10.1038/nrn2533. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Nakajima K, Kohsaka S. Microglia: activation and their significance in the central nervous system. J Biochem(Tokyo) 2001;130:169–175. doi: 10.1093/oxfordjournals.jbchem.a002969. [DOI] [PubMed] [Google Scholar]
  72. Nasu-Tada K, Koizumi S, Tsuda M, Kunifusa E, Inoue K. Possible involvement of increase in spinal fibronectin following peripheral nerve injury in upregulation of microglial P2X4, a key molecule for mechanical allodynia. Glia. 2006;53:769–775. doi: 10.1002/glia.20339. [DOI] [PubMed] [Google Scholar]
  73. Nimmerjahn A, Kirchhoff F, Helmchen F. Resting microglial cells are highly dynamic surveillants of brain parenchyma in vivo. Science. 2005;308:1314–1318. doi: 10.1126/science.1110647. [DOI] [PubMed] [Google Scholar]
  74. North RA. Molecular physiology of P2X receptors. Physiol Rev. 2002;82:1013–1067. doi: 10.1152/physrev.00015.2002. [DOI] [PubMed] [Google Scholar]
  75. Obata K, Yamanaka H, Dai Y, Mizushima T, Fukuoka T, Tokunaga A, Noguchi K. Differential activation of MAPK in injured and uninjured DRG neurons following chronic constriction injury of the sciatic nerve in rats. Eur J Neurosci. 2004;20:2881–2895. doi: 10.1111/j.1460-9568.2004.03754.x. [DOI] [PubMed] [Google Scholar]
  76. Obata N, Mizobuchi S, Itano Y, Matsuoka Y, Kaku R, Tomotsuka N, Morita K, Kanzaki H, Ouchida M, Yokoyama M. Decoy strategy targeting the brain-derived neurotrophic factor exon I to attenuate tactile allodynia in the neuropathic pain model of rats. Biochem Biophys Res Commun. 2011;408:139–144. doi: 10.1016/j.bbrc.2011.03.137. [DOI] [PubMed] [Google Scholar]
  77. Perez-Medrano A, Donnelly-Roberts DL, Honore P, Hsieh GC, Namovic MT, Peddi S, Shuai Q, Wang Y, Faltynek CR, Jarvis MF, Carroll WA. Discovery and biological evaluation of novel cyanoguanidine P2X(7) antagonists with analgesic activity in a rat model of neuropathic pain. J Med Chem. 2009;52:3366–3376. doi: 10.1021/jm8015848. [DOI] [PubMed] [Google Scholar]
  78. Portanova JP, Zhang Y, Anderson GD, Hauser SD, Masferrer JL, Seibert K, Gregory SA, Isakson PC. Selective neutralization of prostaglandin E2 blocks inflammation, hyperalgesia, and interleukin 6 production in vivo. J Exp Med. 1996;184:883–891. doi: 10.1084/jem.184.3.883. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Qureshi OS, Paramasivam A, Yu JC, Murrell-Lagnado RD. Regulation of P2X4 receptors by lysosomal targeting, glycan protection and exocytosis. J Cell Sci. 2007;120:3838–3849. doi: 10.1242/jcs.010348. [DOI] [PubMed] [Google Scholar]
  80. Rao JS, Ertley RN, Lee HJ, DeMar JC, Jr, Arnold JT, Rapoport SI, Bazinet RP. n-3 polyunsaturated fatty acid deprivation in rats decreases frontal cortex BDNF via a p38 MAPK-dependent mechanism. Mol Psychiatry. 2007;12:36–46. doi: 10.1038/sj.mp.4001888. [DOI] [PubMed] [Google Scholar]
  81. Royle SJ, Bobanovic LK, Murrell-Lagnado RD. Identification of a non-canonical tyrosine-based endocytic motif in an ionotropic receptor. J Biol Chem. 2002;277:35378–35385. doi: 10.1074/jbc.M204844200. [DOI] [PubMed] [Google Scholar]
  82. Samad TA, Sapirstein A, Woolf CJ. Prostanoids and pain: unraveling mechanisms and revealing therapeutic targets. Trends Mol Med. 2002;8:390–396. doi: 10.1016/s1471-4914(02)02383-3. [DOI] [PubMed] [Google Scholar]
  83. Sasaki Y, Hoshi M, Akazawa C, Nakamura Y, Tsuzuki H, Inoue K, Kohsaka S. Selective expression of Gi/o-coupled ATP receptor P2Y12 in microglia in rat brain. Glia. 2003;44:242–250. doi: 10.1002/glia.10293. [DOI] [PubMed] [Google Scholar]
  84. Scholz J, Woolf CJ. Can we conquer pain? Nat Neurosci. 2002;5(Suppl):1062–1067. doi: 10.1038/nn942. [DOI] [PubMed] [Google Scholar]
  85. Seil M, El OM, Fontanils U, Etxebarria IG, Pochet S, Dal MG, Marino A, Dehaye JP. Ivermectin-dependent release of IL-1beta in response to ATP by peritoneal macrophages from P2X(7)-KO mice. Purinergic Signal. 2010;6:405–416. doi: 10.1007/s11302-010-9205-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Shinozaki Y, Sumitomo K, Tsuda M, Koizumi S, Inoue K, Torimitsu K. Direct Observation of ATP-Induced Conformational Changes in Single P2X4 Receptors. PLoS Biol. 2009;7:e103. doi: 10.1371/journal.pbio.1000103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Skaper SD, Debetto P, Giusti P. The P2X7 purinergic receptor: from physiology to neurological disorders. FASEB J. 2010;24:337–345. doi: 10.1096/fj.09-138883. [DOI] [PubMed] [Google Scholar]
  88. Sweitzer SM, White KA, Dutta C, DeLeo JA. The differential role of spinal MHC class II and cellular adhesion molecules in peripheral inflammatory versus neuropathic pain in rodents. J Neuroimmunol. 2002;125:82–93. doi: 10.1016/s0165-5728(02)00036-x. [DOI] [PubMed] [Google Scholar]
  89. Tanga FY, Raghavendra V, DeLeo JA. Quantitative real-time RT-PCR assessment of spinal microglial and astrocytic activation markers in a rat model of neuropathic pain. Neurochem Int. 2004;45:397–407. doi: 10.1016/j.neuint.2003.06.002. [DOI] [PubMed] [Google Scholar]
  90. Tao X, Finkbeiner S, Arnold DB, Shaywitz AJ, Greenberg ME. Ca2+ influx regulates BDNF transcription by a CREB family transcription factor-dependent mechanism. Neuron. 1998;20:709–726. doi: 10.1016/s0896-6273(00)81010-7. [DOI] [PubMed] [Google Scholar]
  91. Toulme E, Soto F, Garret M, Boue-Grabot E. Functional properties of internalization-deficient P2X4 receptors reveal a novel mechanism of ligand-gated channel facilitation by ivermectin. Mol Pharmacol. 2006;69:576–587. doi: 10.1124/mol.105.018812. [DOI] [PubMed] [Google Scholar]
  92. Toyomitsu E, Tsuda M, Yamashita T, Tozaki-Saitoh H, Tanaka Y, Inoue K. CCL2 promotes P2X4 receptor trafficking to the cell surface of microglia. Purinergic Signal. 2012 doi: 10.1007/s11302-011-9288-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Tozaki-Saitoh H, Tsuda M, Miyata H, Ueda K, Kohsaka S, Inoue K. P2Y12 receptors in spinal microglia are required for neuropathic pain after peripheral nerve injury. J Neurosci. 2008;28:4949–4956. doi: 10.1523/JNEUROSCI.0323-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Trang T, Beggs S, Salter MW. Purinoceptors in microglia and neuropathic pain. Pflugers Arch. 2006;452:645–652. doi: 10.1007/s00424-006-0074-5. [DOI] [PubMed] [Google Scholar]
  95. Trang T, Beggs S, Salter MW. ATP receptors gate microglia signaling in neuropathic pain. Exp Neurol. 2011 doi: 10.1016/j.expneurol.2011.11.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Trang T, Beggs S, Wan X, Salter MW. P2X4-receptor-mediated synthesis and release of brain-derived neurotrophic factor in microglia is dependent on calcium and p38-mitogen-activated protein kinase activation. J Neurosci. 2009;29:3518–3528. doi: 10.1523/JNEUROSCI.5714-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Tsuda M, Inoue K, Salter MW. Neuropathic pain and spinal microglia: a big problem from molecules in “small” glia. Trends Neurosci. 2005;28:101–107. doi: 10.1016/j.tins.2004.12.002. [DOI] [PubMed] [Google Scholar]
  98. Tsuda M, Kuboyama K, Inoue T, Nagata K, Tozaki-Saitoh H, Inoue K. Behavioral phenotypes of mice lacking purinergic P2X4 receptors in acute and chronic pain assays. Mol Pain. 2009a;5:28. doi: 10.1186/1744-8069-5-28. [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Tsuda M, Masuda T, Kitano J, Shimoyama H, Tozaki-Saitoh H, Inoue K. IFN-gamma receptor signaling mediates spinal microglia activation driving neuropathic pain. Proc Natl Acad Sci USA. 2009b;106:8032–8037. doi: 10.1073/pnas.0810420106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Tsuda M, Mizokoshi A, Shigemoto-Mogami Y, Koizumi S, Inoue K. Activation of p38 mitogen-activated protein kinase in spinal hyperactive microglia contributes to pain hypersensitivity following peripheral nerve injury. Glia. 2004;45:89–95. doi: 10.1002/glia.10308. [DOI] [PubMed] [Google Scholar]
  101. Tsuda M, Shigemoto-Mogami Y, Koizumi S, Mizokoshi A, Kohsaka S, Salter MW, Inoue K. P2X4 receptors induced in spinal microglia gate tactile allodynia after nerve injury. Nature. 2003;424:778–783. doi: 10.1038/nature01786. [DOI] [PubMed] [Google Scholar]
  102. Tsuda M, Toyomitsu E, Komatsu T, Masuda T, Kunifusa E, Nasu-Tada K, Koizumi S, Yamamoto K, Ando J, Inoue K. Fibronectin/integrin system is involved in P2X(4) receptor upregulation in the spinal cord and neuropathic pain after nerve injury. Glia. 2008a;56:579–585. doi: 10.1002/glia.20641. [DOI] [PubMed] [Google Scholar]
  103. Tsuda M, Toyomitsu E, Kometani M, Tozaki-Saitoh H, Inoue K. Mechanisms underlying fibronectin-induced up-regulation of P2X4R expression in microglia: distinct roles of PI3K-Akt and MEK-ERK signalling pathways. J Cell Mol Med. 2009c;13:3251–3259. doi: 10.1111/j.1582-4934.2009.00719.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Tsuda M, Tozaki-Saitoh H, Masuda T, Toyomitsu E, Tezuka T, Yamamoto T, Inoue K. Lyn tyrosine kinase is required for P2X(4) receptor upregulation and neuropathic pain after peripheral nerve injury. Glia. 2008b;56:50–58. doi: 10.1002/glia.20591. [DOI] [PubMed] [Google Scholar]
  105. Ulmann L, Hatcher JP, Hughes JP, Chaumont S, Green PJ, Conquet F, Buell GN, Reeve AJ, Chessell IP, Rassendren F. Up-regulation of P2X4 receptors in spinal microglia after peripheral nerve injury mediates BDNF release and neuropathic pain. J Neurosci. 2008;28:11263–11268. doi: 10.1523/JNEUROSCI.2308-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Ulmann L, Hirbec H, Rassendren F. P2X4 receptors mediate PGE2 release by tissue-resident macrophages and initiate inflammatory pain. EMBO J. 2010;29:2290–2300. doi: 10.1038/emboj.2010.126. [DOI] [PMC free article] [PubMed] [Google Scholar]
  107. Voscopoulos C, Lema M. When does acute pain become chronic? Br J Anaesth. 2010;105(Suppl 1):i69–i85. doi: 10.1093/bja/aeq323. [DOI] [PubMed] [Google Scholar]
  108. Watkins LR, Maier SF. Glia: a novel drug discovery target for clinical pain. Nat Rev Drug Discov. 2003;2:973–985. doi: 10.1038/nrd1251. [DOI] [PubMed] [Google Scholar]
  109. Watkins LR, Milligan ED, Maier SF. Glial activation: a driving force for pathological pain. Trends Neurosci. 2001;24:450–455. doi: 10.1016/s0166-2236(00)01854-3. [DOI] [PubMed] [Google Scholar]
  110. West AE, Chen WG, Dalva MB, Dolmetsch RE, Kornhauser JM, Shaywitz AJ, Takasu MA, Tao X, Greenberg ME. Calcium regulation of neuronal gene expression. Proc Natl Acad Sci USA. 2001;98:11024–11031. doi: 10.1073/pnas.191352298. [DOI] [PMC free article] [PubMed] [Google Scholar]
  111. Woolf CJ, Salter MW. Neuronal plasticity: increasing the gain in pain. Science. 2000;288:1765–1769. doi: 10.1126/science.288.5472.1765. [DOI] [PubMed] [Google Scholar]
  112. Yuan H, Zhu X, Zhou S, Chen Q, Zhu X, Ma X, He X, Tian M, Shi X. Role of mast cell activation in inducing microglial cells to release neurotrophin. J Neurosci Res. 2010;88:1348–1354. doi: 10.1002/jnr.22304. [DOI] [PubMed] [Google Scholar]
  113. Zhang J, De Koninck Y. Spatial and temporal relationship between monocyte chemoattractant protein-1 expression and spinal glial activation following peripheral nerve injury. J Neurochem. 2006;97:772–783. doi: 10.1111/j.1471-4159.2006.03746.x. [DOI] [PubMed] [Google Scholar]
  114. Zhao J, Seereeram A, Nassar MA, Levato A, Pezet S, Hathaway G, Morenilla-Palao C, Stirling C, Fitzgerald M, McMahon SB, Rios M, Wood JN. Nociceptor-derived brain-derived neurotrophic factor regulates acute and inflammatory but not neuropathic pain. Mol Cell Neurosci. 2006 doi: 10.1016/j.mcn.2005.11.008. [DOI] [PubMed] [Google Scholar]
  115. Zhou Z, Hong EJ, Cohen S, Zhao WN, Ho HY, Schmidt L, Chen WG, Lin Y, Savner E, Griffith EC, Hu L, Steen JA, Weitz CJ, Greenberg ME. Brain-specific phosphorylation of MeCP2 regulates activity-dependent Bdnf transcription, dendritic growth, and spine maturation. Neuron. 2006;52:255–269. doi: 10.1016/j.neuron.2006.09.037. [DOI] [PMC free article] [PubMed] [Google Scholar]
  116. Zimmermann M. Pathobiology of neuropathic pain. Eur J Pharmacol. 2001;429:23–37. doi: 10.1016/s0014-2999(01)01303-6. [DOI] [PubMed] [Google Scholar]

RESOURCES