Cell Tissue Res (2014) 357:407–426 DOI 10.1007/s00441-014-1942-5

REVIEW

Neuronal calcium signaling in chronic pain Anna M. Hagenston & Manuela Simonetti

Received: 16 April 2014 / Accepted: 3 June 2014 / Published online: 12 July 2014 # Springer-Verlag Berlin Heidelberg 2014

Abstract Acute physiological pain, the unpleasant sensory response to a noxious stimulus, is essential for animals and humans to avoid potential injury. Pathological pain that persists after the original insult or injury has subsided, however, not only results in individual suffering but also imposes a significant cost on society. Improving treatments for longlasting pathological pain requires a comprehensive understanding of the biological mechanisms underlying pain perception and the development of pain chronicity. In this review, we aim to highlight some of the major findings related to the involvement of neuronal calcium signaling in the processes that mediate chronic pain. Keywords Sensitization . Chronic pain . Pain pathways . Calcium signaling . Nuclear calcium

Introduction The sensation of pain is important for the survival of the individual. Pain sensation is propagated from the periphery into the central nervous system via pain pathways originating in primary sensory neurons whose cell bodies are situated within the dorsal root ganglia (DRG). These neurons have terminal endings within the skin and viscera, where noxious stimuli from the environment are detected, and carry this information via unmyelinated C fibers and myelinated Aδ fibers to the spinal cord dorsal horn, where they form A. M. Hagenston (*) University of Heidelberg, Neurobiology, Im Neuenheimer Feld 364, 69120 Heidelberg, Germany e-mail: [email protected] M. Simonetti University of Heidelberg, Pharmacology, Im Neuenheimer Feld 364, 69120 Heidelberg, Germany

excitatory synapses with neurons in laminae I, II, and V. Second order neurons in the spinal cord dorsal horn project in turn via the spinothalamic and spinoreticulothalamic relays to the somatosensory and limbic systems of the cortex, respectively, where the nociceptive and emotional/aversive components of pain are processed (Millan 1999). Together with these ascending pain pathways that convey pain stimuli from the periphery to higher brain centers, there are also painmodulating descending pathways. These arise in the brainstem, hypothalamus, and cortical structures, and modulate both sensory inputs from primary afferent fibers and projection neurons in the dorsal horn of the spinal cord (Millan 1999, 2002). The best known of these pathways are the axes of the periaqueductal gray and rostroventral medulla, which can both inhibit or facilitate sensory processing in the spinal cord dorsal horn. Just as the sensation of pain involves the activity of a neuronal network spanning from the periphery to the spinal cord to the brain and back, so too the development of pain hypersensitivity involves changes at all levels of this network, beginning with heightened sensitivity of peripheral neurons (peripheral sensitization), altered pain processing within the spinal cord and deriving from descending projections (central sensitization), and finally via changes in the experience of pain following from altered cerebral network activity (also central sensitization). Sensitization of the neuronal networks underlying pain perception results behaviorally in spontaneously occurring pain, increased responsiveness to noxious stimuli (hyperalgesia), and nociceptive responsiveness to otherwise innocuous stimuli (allodynia). In the periphery, this sensitization is characterized by an increase in the basal firing rates of primary sensory neurons, their elevated responsiveness to noxious stimuli, and a decrease in their activation thresholds for thermal and mechanical stimulation (Woolf and Ma 2007). Central sensitization at the level of the spinal cord is considered to result from activity-dependent changes in spinal neuronal function, and involves both the long-term

408

potentiation (LTP) of individual synapses from peripheral afferents onto spinal neurons as well as the increased excitability of neurons within the spinal cord dorsal horn (Ji et al. 2003; Sandkühler 2009). Within the somatosensory and limbic systems of the brain, central sensitization takes the form of heightened neuronal responsiveness and neural circuit reorganization (Moseley and Flor 2012). The neuronal plasticity associated with peripheral and central sensitization manifests itself in a range of different functional and structural plastic changes that together lead to longlasting alterations in peripheral sensitivity and spinal and cortical neuronal networks, an elevated net spinal output, an altered perception of pain, and the development of a pathological pain state. Functionally, neuronal plasticity can occur on at least four levels, including (1) changes in neuronal excitability and signal transduction resulting from altered protein expression of and/or post-translational modifications to existent proteins; (2) changes in synaptic strength deriving postsynaptically from changes in the number and/or synaptic localization of neurotransmitter receptors and/or presynaptically from changes in neurotransmitter release probability; (3) changes in synaptic connectivity resulting from modifications in the density and/or size of synaptic spines, altered complexity of neuronal dendritic arbors and/or the numbers of axonal afferents; and (4) changes in the numbers of cells that collectively respond to a given noxious stimulus. Our aim here is to highlight the role of calcium signaling in some of these neuronal plastic processes. Numerous diseases of the central nervous system, including pathological pain, are associated with abnormal neuronal calcium homeostasis and calcium signaling (Fernyhough and Calcutt 2010; Mattson 2012; Stutzmann 2007). Indeed, aberrant calcium channel physiology and expression has been implicated in not only neuropathic pain and diabetic neuropathy but also chronic inflammatory pain (Bourinet et al. 2014; Fernyhough and Calcutt 2010; Naziroglu et al. 2012). Nevertheless, a number of marked differences in the underlying mechanisms of the different pain modalities have been identified, and results from studies examining calcium regulation in a variety of different pain models suggest the existence of specific patterns of injury and/or disease-induced changes in the regulation of calcium (Xu and Yaksh 2011). Regardless of the particular mechanisms involved, the net output in all of the different forms of chronic pain is an over-activation of pain-transducing primary sensory neurons in the DRG and an enhanced excitation of second order neurons in the spinal cord dorsal horn.

Neuropathic pain Neuropathic pain is defined as pain whose initiating cause is a primary lesion or dysfunction in the nervous system. In the

Cell Tissue Res (2014) 357:407–426

spinal cord, neuropathic pain is characterized by the increased responsiveness and spontaneous activity of dorsal horn neurons, the expansion of the receptive fields of these neurons, and a loss of GABAergic inhibition (Campbell and Meyer 2006; Kuner 2010; von Hehn et al. 2012). Many of the mechanisms proposed to explain the primary genesis of neuropathic pain, however, involve the increased excitability or enhanced spontaneous activity of primary afferents. Such elevated activity may result from a number of possible functional alterations in the afferent cells, including the increased sensitivity and/or activity of peripheral terminals, cell bodies, and afferent axons of injured neurons, the formation by injured cells of ectopic neuronal pacemakers along sensory nerve fibers, and the increased electrical excitability of adjacent, uninjured axons (Campbell and Meyer 2006; von Hehn et al. 2012). It is well established, for instance, that nerve injury in general and axotomy in particular leads to increased neuronal excitability of DRG neurons (Study and Kral 1996; Wall and Devor 1983). This increased excitability is strongly associated with decreases in voltage-dependent calcium currents, a decrease in resting calcium concentrations in injured DRG neurons, a decrease in the magnitude of evoked calcium transients specifically in injured putative nociceptive DRG neurons, and finally a diminished recruitment of calciumsensitive potassium channels (Fuchs et al. 2005, 2007; Hogan 2007).

Diabetic neuropathy Diabetic neuropathy is one of the most common peripheral neuropathies. In patients, diabetic neuropathy leads to the degeneration of peripheral nerve fibers and a concomitant loss of sensory perception. One of the foremost symptoms of diabetic neuropathy is pain. Both the painful and degenerative aspects of diabetic neuropathy have been linked to abnormalities in calcium signaling and homeostasis. Indeed, disturbed calcium homeostasis has been observed in a variety of tissues from animals with experimentally-induced diabetes as well as from diabetic patients, and in sensory neurons typically consists of an increase in the resting intracellular calcium concentration, the decreased activity of calcium transporters, and a reduction in stimulus-triggered calcium rises (Fernyhough and Calcutt 2010; Verkhratsky and Fernyhough 2008). One source of these changes is a decrease in the rate of calcium extrusion following stimulation in diabetic neurons, which can result from a decrease in the expression of plasma membrane calcium pumps as well as a reduction in the calcium uptake by the neuronal endoplasmic reticulum (ER) and mitochondria (Fernyhough and Calcutt 2010; Verkhratsky and Fernyhough 2008; Zherebitskaya et al. 2012). Another major source of calcium dysregulation in diabetic neuropathy is a change in the entry of calcium through plasma membrane

Cell Tissue Res (2014) 357:407–426

voltage-dependent calcium channels (VDCCs) (Todorovic and Jevtovic-Todorovic 2014). In rodent models of diabetes, for example, enhanced mRNA levels of and currents from Ttype VDCCs have been reported in putative nociceptive DRG neurons, and these changes are correlated with the development of diabetic pain (Hall et al. 1995; Jagodic et al. 2007; Khomula et al. 2013; Yusaf et al. 2001). Moreover, both the pharmacological inhibition and knockdown of T-type channels was able to reduce pain in animal models of diabetes (Latham et al. 2009; Messinger et al. 2009; Obradovic et al. 2014). These results, and others, suggest that increased expression of T-type VDCC channels may contribute to the underlying mechanisms of painful diabetic neuropathy (Todorovic and Jevtovic-Todorovic 2014). A particularly interesting molecular mechanism for controlling the expression and function of T-type VDCC subunits under hyperglycemic conditions—as independently shown in two recent papers—involves their asparagine-linked glycosylation (Orestes et al. 2013; Weiss et al. 2013). Despite some differences, possibly due to different cell culture conditions, together these studies convincingly demonstrate that pathological glucose levels in a recombinant cell culture system in vitro and in the leptin-deficient mouse model of peripheral diabetic neuropathy in vivo lead to the glycosylation of extracellular asparagine residues of CaV3.2 subunits, and that this glycosylation results in the augmented surface expression, increased current density, and accelerated current kinetics of T-type VDCCs (Orestes et al. 2013; Weiss et al. 2013). These findings raise the exciting possibility that pharmacological agents directed at glycosylation sites on T-type VDCCs may provide a significant therapeutic benefit to sufferers of painful diabetic neuropathy. Indeed, peripheral application of a glycoside hydrolase both inhibited native T-type currents and reversed hyperalgesic responses in leptin-deficient diabetic mice (Orestes et al. 2013). Although these two studies focused their efforts on the examination of CaV3.2, the most prevalent T-type VDCC isoform in small DRG cells, all three isoforms of T-type VDCCs have been observed in DRG neurons (Choi et al. 2007; Talley et al. 1999). Given the high degree of conservation of asparagine glycosylation sites between Ttype VDCC isoforms, it is therefore conceivable that glycosylation may also modulate the expression and function of CaV3.1 and CaV3.3 T-type VDCC isoforms (Orestes et al. 2013; Weiss et al. 2013). In light of this, it is interesting to note that at least two recent studies have suggested that the modulation of T-type VDCC-mediated currents in DRG neurons may involve changes not only in the surface expression but also in the isoform composition of these channels (Khomula et al. 2014; Yue et al. 2013). In particular, specific types of diabetic neuropathy were associated with the differential remodeling of T-type VDCC isoforms which have distinct biophysical properties, such that the development of thermal hyperalgesia

409

specifically (versus hypo- or normalgesia) involved an increase in the peak current density associated with fast inactivating T-type current and a lack of slowly inactivating T-type currents in a subset of nonmyelinated DRG neurons (Khomula et al. 2014). While fast inactivating T-type currents were mediated predominantly by CaV3.2 subunits, slow inactivating T-type currents were mediated by CaV3.2 and other T-type VDCC subunits. The lack of slow T-type currents in diabetic thermal hyperalgesia thus suggests an elimination either of neurons expressing this current or of the underlying subunits (Weiss et al. 2013). Chronic inflammatory pain When peripheral tissues are chronically inflamed, primary sensory nerve terminals are surrounded by a diversity of chemical mediators that stimulate and sensitize DRG neurons, resulting in a number of long-lasting functional changes (Basbaum et al. 2009). Two prominent changes in DRG neurons in models of chronic inflammatory pain are increases in the basal concentration of cytosolic calcium ions and altered voltage-dependent calcium currents (Lu and Gold 2008; Lu et al. 2010; Waxman and Zamponi 2014). Resting calcium levels in medium diameter sensory neurons, for example, were seen to be higher, and depolarization-evoked calcium transients to be larger and to decay more slowly in small and medium diameter sensory neurons from inflamed tissues (Lu and Gold 2008). Consistent with an increase in the expression and/or permeability of plasma membrane ion channels, the altered calcium regulation of sensory neurons has been observed to be accompanied by a notable elevation of their excitability and resting membrane conductance (Dang et al. 2005; Flake et al. 2005; Moore et al. 2002). Indeed, the mechanisms underlying peripheral and central sensitization during inflammation involve the increased surface trafficking of sodium- and calcium-permeable inflammatory receptors and voltage-gated ion channels in primary sensory neurons (e.g., ASIC1, TRPV1, TRPA1, P2X3, NaV1.8, and N-type VDCCs), but also changes in the trafficking and composition of synaptic and extrasynaptic glutamate receptors (reviewed in Bourinet et al. 2014; Ma and Quirion 2014; see below). Interestingly, however, the increased excitability of sensory neurons during inflammation seems not to involve upregulated calcium influx at the soma, but is thought instead to specifically involve changes in calcium signaling within sensory terminals and presynaptic endings (Basbaum et al. 2009; Katz and Gold 2006). Accordingly, the contribution of increased calcium levels in DRG cells to inflammatory hyperalgesia is most likely to result from the facilitation of transmitter release both in the periphery and at central terminals, and the consequent enhancement of both neurogenic inflammation and nociceptive neurotransmission, respectively.

410

Calcium sources and sinks Calcium is an integral second messenger in neurons, and is involved in the functional and structural neuronal plasticity that underlies peripheral and central sensitization at all levels. Changes in the intracellular concentration of calcium ions in peripheral and central neurons are mediated by calcium entry through ligand-gated channels such as N-methyl-D-aspartate receptors (NMDARs), calcium-permeable α-amino-3-hydroxyl-5-methyl-4-isoxazolepropionate receptors (AMPARs), and ATP-responsive purinoreceptors (P2XRs) that respond to synaptically released neurotransmitter and glial- or microglial-derived ATP, respectively; through acidsensing ion channels (ASICs), which are activated by extracellular protons; through transient receptor potential (TRP) channels such as the heat- and capsaicin-responsive vallinoid (TRPV) receptors; through VDCCs, including both highvoltage-activated (HVA) and low-voltage-activated (LVA) subtypes that are activated in response to neuronal depolarization (see Bourinet et al. 2014 for a comprehensive overview of NMDARs, P2XRs, TRPs, and ASICs in afferent pain pathways and their roles in pain pathophysiology). Changes in the concentration of calcium ions in neurons along pain pathways are also mediated by the release of calcium from intracellular stores following activation of G-protein-coupled receptors (GPCRs) or receptor tyrosine kinases (RTKs) by their respective ligands and subsequent stimulation of ER inositol trisphosphate receptors (IP3Rs); by the transport out of the mitochondrial lumen via sodium/calcium/lithium exchangers (NCLX); and by the intracellular release of calcium following the calcium-dependent activation of ER ryanodine receptors (RyRs). Although calcium sources typically represent the more popular subjects for study, calcium sinks and extrusion mechanisms, too, are important modulators of calcium signaling, and have also been implicated in the neuronal mechanisms underlying pain sensitization. Neuronal calcium sinks include plasma membrane calcium ATPases (PMCA) and sodium/calcium exchangers (NCXs), which remove cytosolic calcium into the extracellular space; sarco/endoplasmic reticulum calcium ATPases (SERCA pumps), which pump calcium ions into the ER; nuclear NCXs, which remove nuclear calcium into the nucleoplasmic reticulum; the mitochondrial uniporter (MCU), which mediates mitochondrial calcium uptake; and a wide range of calcium binding proteins (CaBPs), including both calcium buffers like calbindin D28k and signaling molecules such as calmodulin (CaM) (Fig. 1).

NMDARs and AMPARs The activation of NMDARs and AMPARs is considered to be a key trigger for activity-dependent changes in neuronal

Cell Tissue Res (2014) 357:407–426

excitability and function. NMDARs and AMPARs are expressed at excitatory postsynaptic sites throughout the spinal cord and supraspinal central nervous system, but also in presynaptic primary sensory neurons and nerve terminals (Millan 1999; Polgar et al. 2008; Wu et al. 2004). Consistent with their involvement in the mechanisms underlying peripheral and central sensitization, the altered expression and/or function of both AMPARs and NMDARs has been observed in altered pain states in primary sensory neurons, in the spinal cord dorsal horn, and in supraspinal regions important for pain sensation and pain perception (Basbaum et al. 2009; Ho et al. 2013; Liu and Salter 2010; Neugebauer et al. 2009; Zhou et al. 2014; Zhuo et al. 2011; and others). NMDARs draw particular attention in the context of neuronal plasticity and pain sensitization due to two key features. First, on account of their calcium permeability, NMDARs represent a major source of synaptic activity-dependent calcium rises. Indeed, following their involvement in regulating a broad range of kinases, phosphatases, and other enzymes, NMDAR-mediated calcium rises have been implicated in nearly all aspects of pain plasticity (Woolf and Salter 2000; Wu and Zhuo 2009; Zhuo et al. 2011), and NMDAR antagonists represent an attractive target in chronic pain therapy (Chizh and Headley 2005; Fisher et al. 2000; Wu and Zhuo 2009). Second, the combined sensitivity of NMDARs to both synaptically released neurotransmitter and membrane potential via the voltage-dependent blockade of the ion-conducting pore by extracellular magnesium ions endows them with the unique ability to act as coincidence detectors for simultaneous presynaptic glutamate release and postsynaptic action potential generation. Indeed, modulation of extracellular magnesium concentrations—and thereby of NMDAR activity and NMDAR-mediated calcium influx—has been identified both as a potential source of pain hypersensitivity and as a potential target for pain therapy (Begon et al. 2000; Felsby et al. 1996; Rondon et al. 2010). Although NMDARs likely represent the most recognized postsynaptic source of neuronal calcium entry in neurons, AMPARs, depending on their subunit composition, can also contribute to postsynaptic calcium rises and calciumdependent plasticity. In particular, AMPARs lacking the GluA2 (formerly called GluR2) subunit are conferred with a much increased calcium permeability and modified current rectification and total channel conductance (Hollmann et al. 1991). A number of lines of evidence implicate AMPARs in general, and calcium-permeable AMPARs in particular, in the signaling mechanisms underlying pain chronicity. It has been shown, for instance, that peripherally located AMPARs are at least partially responsible for the transmission of acute nociceptive stimuli from muscle fibers and for the development of irritant-induced nocifensive behaviors (Chun et al. 2008). Similarly, the deletion of calcium-permeable AMPARs from peripheral, pain-sensing neurons results in decreased hypersensitivity and sensitization in models of chronic

Cell Tissue Res (2014) 357:407–426

411

Fig. 1 Calcium sources and sinks. Calcium can enter the cytosol of neurons along the pain axis from the extracellular space via NMDARs and calcium-permeable AMPARs, ATP-responsive P2XRs, ASICs, TRP channels, and VDCCs. Calcium can also enter the cytosol when it is released from intracellular calcium stores via RyRs, or via IP3Rs following the activation of Gq/11-coupled GPCRs or RTKs on the plasma membrane. Calcium can enter the cytosol from the mitochondrial matrix via NCLX, and can be taken up into mitochondria via the MCU. Calcium can also be removed from the cytosol when it is taken up into the ER via SERCA pumps and nuclear NCX, bound by cytosolic CaBPs, or extruded into the extracellular space via PMCAs or NCXs. Cytosolic calcium can freely diffuse through nuclear pores into the nucleus, where it regulates the activity of a wide range of transcription factors (TFs) and controls the expression of diverse target genes

inflammatory pain and arthritis (Gangadharan et al. 2011). Further, the specific antagonism of GluA2-lacking AMPARs was found to alleviate inflammatory pain without affecting animal responses to physiological painful stimuli, indicating an important pathophysiological role for peripheral calciumpermeable AMPARs in chronic pain states (Gangadharan et al. 2011). Indeed, while mice lacking calcium-permeable AMPARs exhibit both a loss of nociceptive plasticity in vitro and a reduction in inflammatory hyperalgesia in vivo, an increase in calcium-permeable AMPARs in GluA2-deficient mice facilitates spinal cord LTP, nociceptive plasticity, and inflammatory hyperalgesia (Hartmann et al. 2004; Youn et al. 2008). Moreover, mice that do not possess calcium-permeable AMPARs lose potentiation and spatial spread of spinal calcium transients induced by peripheral inflammation, indicating that spinal AMPARs are crucial for activity-dependent changes in synaptic processing of nociceptive inputs (Luo et al. 2008). Interestingly, neurons of the spinal cord dorsal horn not only express an unusually high density of calciumpermeable AMPARs (Gu et al. 1996; Tong and MacDermott 2006) but their expression levels relative to calciumimpermeable AMPARs are increased in animal models of both persistent inflammatory and neuropathic pain (Chen et al. 2013; Park et al. 2009; Vikman et al. 2008). These expression changes depend on NMDAR activation, the calcium-dependent internalization of GluA2 AMPAR

subunits, and the insertion of GluR1 AMPAR subunits (Chen et al. 2013; Larsson and Broman 2008; Park et al. 2009). Importantly, increased expression of calciumpermeable AMPARs at postsynaptic sites, for example following the calcium/CaM kinase II (CaMKII)-dependent recruitment of the AMPAR-interacting protein GRIP or the activation of protein kinase C (PKC), both enhances postsynaptic excitation and postsynaptic calcium influx and strengthens the AMPAR-mediated component of spinal synaptic transmission (Galan et al. 2004; Gu et al. 1996; Hartmann et al. 2004; Kopach et al. 2013; Park et al. 2009). Calcium entry through AMPARs may also contribute to the down-regulation of inhibitory glycinergic synaptic function in spinal cord neurons via a mechanism involving CaMKII and calcineurin (Xu et al. 1999). Finally, GluR1-, but not GluA2containing AMPARs in the anterior cingulate cortex, a cortical brain region that, together with the somatosensory cortex and amygdala, plays an important role in pain perception and central sensitization, are implicated in the stimulation of MAP kinases and the induction of LTP in inflammatory pain models in vivo, and their expression levels are elevated in a rat model of visceral hypersensitivity (Toyoda et al. 2009; Zhou et al. 2014). Importantly, the excessive activation of glutamate receptors in general, and NMDARs in particular, is strongly implicated in the mechanisms underlying excitotoxicity in neurons, and

412

cell death has been reported in the central nervous system in models of neuropathic pain (Hardingham 2009; Hardingham and Bading 2010; Leong et al. 2011). It is therefore particularly interesting that, in models of peripheral inflammation and neuropathic pain, which lead to increased spontaneous and evoked activity of spinal nociceptive afferents, neurons may express a mechanism by which NMDARs can regulate their own activity in a negative feedback loop (Tappe et al. 2006). A principal player in this feedback loop is Homer1, an adapter protein that in neurons is concentrated in dendritic spines and at the postsynaptic density (Jardin et al. 2013; Xiao et al. 2000). Interestingly, whereas the Homer1b and Homer1c variants link G-protein-coupled metabotropic glutamate receptors (mGluRs) to internal calcium release channels (IP3Rs) at the ER, and to VDCCs and components of the NMDAR complex at the cell surface, the short, activitydependent splice variant Homer1a antagonizes these interactions (Ango et al. 2001; Jardin et al. 2013; Sala et al. 2003). Consistent with the possibility that Homer1a may regulate synaptic NMDAR and AMPAR activity, cells in the hippocampus of Homer1a knockout animals exhibited larger AMPA/NMDA current ratios than wild-type animals, and a redistribution of GluA2 subunits from the dendritic compartment to the soma. Cells in the hippocampus of transgenic Homer1a over-expressing animals, in contrast, exhibited smaller postsynaptic AMPA/NMDA current ratios, an increase in the proportion of GluA2-containing, calciumimpermeable AMPARs at the synapse, and an abolishment of the maintenance phase of LTP (Rozov et al. 2012). In the spinal cord dorsal horn, up-regulation of Homer1a was observed following chronic constriction injury and peripheral inflammation, and this up-regulation depended on the activation of NMDARs and the subsequent stimulation of the MAP kinase pathway (Miyabe et al. 2006; Tappe et al. 2006). Whereas prevention of Homer1a up-regulation using shRNA exacerbated the inflammatory pain phenotype, viral-mediated over-expression of Homer1a attenuated inflammatory hypersensitivity (Tappe et al. 2006). In another series of experiments, knockout of Homer1a exacerbated, and viral-mediated over-expression of Homer1a in the spinal cord dorsal horn attenuated, neuropathic pain hypersensitivity following chronic constriction injury (Obara et al. 2013). Consistent with these findings are the results of a pharmacological study wherein chronic intraspinal application of [−]-huperzine A, a potent and reversible acetylcholinesterase and NMDAR antagonist, prevented Homer1a up-regulation and retained GluA2-containing, calcium-impermeable AMPARs at synaptic sites following chronic constriction injury (Yu et al. 2013). On a behavioral level, [−]-huperzine A treatment resulted in a diminished pain phenotype (Yu et al. 2013). Thus, the NMDAR-mediated up-regulation of Homer1a in the spinal cord may represent an intrinsic mechanism by which

Cell Tissue Res (2014) 357:407–426

neurons can curb their excitability and calcium responsiveness, and poses an interesting possible target for future pain therapies.

P2XRs P2X purinoreceptors are ligand-gated, ATP-sensitive, nonselective cation channels. Their activation leads to channel opening, and can result in the downstream stimulation of VDCCs and calcium-dependent MAP kinases (Cruz and Cruz 2007; Kaczmarek-Hajek et al. 2012). In particular, peripheral sensory neurons, but also neurons within the spinal cord dorsal horn, express P2XRs, including the P2X1, P2X2, P2X3, and P2X4 subtypes, and exhibit increased excitability in response to stimulation by ATP (Kim et al. 2003; Kuroda et al. 2012). Indeed, ATP not only functions as a pronociceptive stimulus by activating peripheral nociceptors but it is also released by sensory neurons at synapses in the spinal cord where its stimulatory action ultimately leads to an increased net spinal cord output (Jarvis 2010). A link between the development of peripheral and central sensitization in inflammatory and neuropathic pain models and the increased activity and expression of P2XRs is well established (DonnellyRoberts et al. 2008; Fabbretti 2013). Intracellular calcium rises are likely to be an important trigger for the regulation of P2XRs. P2X 3Rs, for instance, are sensitized by PKCdependent signaling pathways downstream from nerve growth factor (NGF) receptors (D’Arco et al. 2007; Giniatullin et al. 2008) and by the calcium/CaM-dependent serine protein kinase, CASK (Gnanasekaran et al. 2013). Moreover, brain-derived neurotrophic factor (BDNF) signaling through calcium, CaMKII, and cyclic AMP response element binding protein (CREB) has been observed to upregulate the expression of P2X3Rs in trigeminal sensory neurons (Simonetti et al. 2008). Interestingly, the co-activation by ATP of P2Y GPCRs, which can trigger the release of calcium from internal stores, may result in reduced P2XR responses due to a loss of plasma membrane phosphoinositides, suggesting that ATP acting on peripheral sensory neurons may simultaneously evoke algesic and analgesic responses (Bernier et al. 2013; Gerevich et al. 2007; Song and Varner 2009). The potential contributions of P2XRs to neuronal pain signaling are two-fold. On the one hand, P2XR activation, which can be triggered by concentrations of ATP down to the nanomolar range, results in neuronal depolarization and increased excitability, and provides a well-documented means by which ATP can activate neurons in pain pathways (Fabbretti 2013; Jarvis 2010). On the other hand, P2XR-derived calcium signals may contribute to the mechanisms involved in pain sensitization, for example by reducing GABA receptor-dependent inhibition (Shrivastava et al. 2011; Sokolova et al. 2001).

Cell Tissue Res (2014) 357:407–426

VDCCs While it is generally accepted that VDCCs regulate fast neurotransmission and neuronal excitability in peripheral and central pain pathways, the specific channel subtypes involved, and the mechanisms by which this regulation occurs, vary across both cell types and pain modalities (Bourinet et al. 2014). LVA (T-type) VDCCs, for example, play a demonstrated role in peripheral as well as central and supraspinal pain processing and sensitization (Bourinet et al. 2005; Nelson and Todorovic 2006). T-type VDCCs are activated near the resting membrane potential and evoke low-threshold calciumdependent spikes or action potentials that can, in turn, trigger bursts of action potentials. Consequently, the activation of Ttype channels on neuronal dendrites is an efficient signal for the transduction of painful stimuli (Cheong and Shin 2013; Todorovic and Jevtovic-Todorovic 2011). Indeed, both the conduction of painful stimuli in peripheral sensory nerves and the gating of sensory information at the level of the thalamus have been reported to depend on T-type VDCC activity (Bourinet et al. 2005; Cheong et al. 2008; Cheong and Shin 2013; Todorovic and Jevtovic-Todorovic 2011). However, while T-type channels in the periphery seem to be involved in pronociceptive functions, in the thalamus they may exert anti-nociceptive effects (Todorovic and Jevtovic-Todorovic 2011). Moreover, although the expression of these channels following chronic constriction injury is upregulated in the spinal cord, their activity is nearly completely lost in sensory fibers (McCallum et al. 2003; Wen et al. 2006). Thus, while the intrathecal application of T-type VDCC antagonists or the down-regulation of their expression in the spinal cord using antisense oligonucleotides was observed to relieve both allodynia and hyperalgesia following chronic constriction injury (Wen et al. 2006), this pain model was associated in sensory neurons with a nearly 60 % reduction of pharmacologically isolated T-type currents and an 80 % reduction in total calcium influx (McCallum et al. 2003). These results support the idea that, in neuropathic pain, the up-regulation of T-type VDCCs is important in the mechanisms of spinal central sensitization, while their downregulation may be involved in peripheral sensitization (Hogan 2007; McCallum et al. 2003; Wen et al. 2010). HVA (L-, N-, and P/Q-type) VDCCs undergo two major changes in peripheral sensory neurons in chronic pain models. On the one hand, peripheral inflammation and spinal nerve ligation have both been observed to trigger a decrease in HVA calcium currents in these cells (Fuchs et al. 2007; Lu et al. 2010; McCallum et al. 2011; Rycroft et al. 2007). This loss of whole-cell calcium currents leads to the decreased stimulation of small (SK) and large (BK) conductance calcium-activated potassium channels, a reduction in action potential-evoked afterhyperpolarizations, and a consequent increase in neuronal excitability (Bahia et al. 2005; Lirk et al. 2008; McCallum

413

et al. 2006; Sarantopoulos et al. 2007; Zhang et al. 2012). In particular, the loss of VDCC-mediated calcium currents is associated with a broadening of action potentials, an elevated maximal firing rate, and a higher incidence of burst firing (Amir and Devor 1997; Hogan 2007; McCallum et al. 2006). Both action potential duration and burst firing were demonstrated to be key determinants of neurotransmitter release and synaptic strength (Lisman 1997; Sabatini and Regehr 1997). Therefore, a decrease in VDCC-mediated calcium currents in nociceptive fibers represents a proximal cause for the increased neurotransmitter release onto and excitability of second order neurons within the spinal cord dorsal horn (Hogan 2007). Supporting this idea are observations that aberrant action potential waveforms and burst firing are directly triggered by antagonism of VDCCs or intracellular calcium buffering, and that the hyperexcitability of nociceptive afferents in chronic pain models can be corrected by a restoration of calcium influx (Hogan et al. 2008; Hogan 2007; Lirk et al. 2008). On the other hand, painful nerve injury is accompanied by an increase in the expression of the α2δ1 auxiliary subunit of HVA VDCCs (Dolphin 2013), and spinal nerve ligation and acute inflammation have both been associated with the increased expression of N-type VDCC protein in afferent sensory neurons and at synaptic locations in the spinal cord dorsal horn (Bourinet et al. 2014; Cizkova et al. 2002; Lu et al. 2010; Yokoyama et al. 2003). HVA VDCCs are multimeric protein complexes formed by the assembly of a pore-forming (α1) subunit together with two auxiliary (β and α2δ) subunits. The α2δ subunit is linked to the plasma membrane via a glycophosphatidylinositol anchor, and is complexed with the extracellular face of the pore-forming subunit (Bourinet et al. 2014). Importantly, α2δ subunits participate neither in pore formation nor in the gating of VDCCs. Rather, inclusion of this subunit in HVA VDCCs seems to have regulatory functions: that of a chaperone which aids in the trafficking of HVA VDCCs to synaptic membranes, and that of a potentiator in VDCC-exocytosis coupling (Davies et al. 2007; Hoppa et al. 2012). Consequently, the increased expression levels of α2δ1 observed in chronic pain models is consistent both with the documented increased trafficking of N-type VDCCs to presynaptic sites and with increased presynaptic neurotransmitter release probability and postsynaptic excitation (Bauer et al. 2009; Bourinet et al. 2014; Hoppa et al. 2012; Li et al. 2014a; Lu et al. 2010). In fact, the antinociceptive action of the drug gabapentin, which binds to α2δ subunits, is thought to derive from its ability to disrupt VDCC trafficking to presynaptic sites (Davies et al. 2007; Hendrich et al. 2008). Consistent with the idea that the up-regulation of α2δ1 may be a proximal cause for pain sensitization is the observation that the exogenous over-expression of α2δ1 results in enhanced VDCC activity in sensory neurons and hyperexcitability in their targets within the spinal cord dorsal horn (Li et al. 2006).

414

Moreover, α2δ1 knockout animals exhibit significantly delayed development of mechanical hypersensitivity in a nerve injury model of neuropathic pain (Patel et al. 2013). Further, blockade of injury-induced spinal nerve activity diminished increases of α2δ1 expression in the DRG, and abolished its up-regulation at synaptic sites in the spinal cord dorsal horn. These changes were correlated with a delayed onset of allodynia following injury (Boroujerdi et al. 2008). Along the same lines, knockdown α2δ1 expression by antisense oligonucleotides reverses allodynia in spinal cord injury and spinal nerve ligation models of neuropathic pain, and in a chronic constriction injury model in the trigeminal nerve (Boroujerdi et al. 2008, 2011; Li et al. 2004, 2014a).

RTKs and GPCRs One major source of calcium in neurons, besides the channels that allow calcium entry from the extracellular space, is the ER. Internal calcium release from the ER into the cytosol is triggered by the activation of phospholipase C (PLC)-linked RTKs and Gq/11-interacting GPCRs that are coupled via the activation of PLC to the production of diacylglycerol (DAG), which together with calcium stimulates protein kinase C (PKC), and inositol trisphosphate (IP3), which triggers the subsequent IP3R-dependent intracellular release of calcium (Berridge 1998; Rhee 2001). Plasma membrane receptors that can evoke the release of calcium in neurons along pain pathways include the NGF and BDNF RTKs, TrkA and TrkB, respectively, P2Y purinoreceptors, group I/II mGluRs, serotonergic (5-HT) receptors, noradrenergic receptors, opioid receptors, and many others (McKelvey et al. 2013; Stone and Molliver 2009; Trang et al. 2011). RTKs and GPCRs, and the calcium rises and calcium-dependent signaling pathways they activate, stimulate a broad range of ion channels in peripheral and central pain pathways, including calcium-permeable acidsensing ion channels, TRP channels, NMDARs, and P2XRs, but also calcium-sensitive chloride and potassium channels (Jin et al. 2013; Kweon and Suh 2013; McKelvey et al. 2013; Schicker et al. 2010; Stone and Molliver 2009). Indeed, the molecular and cellular consequences of RTK and GPCR activation and internal calcium release are established mediators of peripheral and central sensitization, and as such represent attractive candidates for potential pain therapies (Liu and Salter 2010; McKelvey et al. 2013). P2Y receptor activity on peripheral sensory neurons, for example, can sensitize TRPV1 channels and inhibit currents through P2XRs and N-type VDCCs (Borvendeg et al. 2003; Gerevich et al. 2005; Moriyama et al. 2003), and is linked to increases in the phosphorylation of MAP kinases and to elevated neuronal excitability deriving from a decrease in potassium channel expression (Li et al. 2014b). Moreover, the activity of P2YRs and the calcium rises they evoke have been shown to

Cell Tissue Res (2014) 357:407–426

make an important contribution to the development of neuropathic and inflammatory pain sensitization (Li et al. 2014b; Malin and Molliver 2010). Pain-modulating descending pathways represent a particularly interesting source of GPCR-stimulating neurotramitters. There exist both inhibiting and facilitating descending pathways, but the serotonergic/noradrenergic pathway and the opioidergic pathway represent the best-characterized sources of descending pain modulation. The dual modality of these descending pathways is thought to be due to the existence of two distinct populations of neurons in the brainstem, “on cells” and “off cells,” which are differentially recruited by higher brain structures in conditions of chronic pain, stress, or fear to facilitate or inhibit pain at the spinal level (Heinricher et al. 2009). Adding to this complexity is the fact that both inhibitory (Gi/o-coupled) and excitatory (Gq/11coupled) GPCRs for descending modulatory neurotransmitters are expressed in the spinal cord dorsal horn (Millan 2002). Thus, the consequences of neuromodulator release will clearly depend on the receptor subtype that is activated. It follows that the relative expression levels of the different receptor classes and the types of neurotransmitter released by each fiber modality may determine the consequences of a given neuromodulatory input. Moreover, nociception-associated changes in relative receptor expression levels thus represent an attractive mechanism for changed responses to descending modulatory inputs. GABAergic interneurons of the spinal cord dorsal horn, for instance, were recently shown to express a much greater proportion of Gq/11-coupled (5-HT1A) than of Gi/o-coupled (5-HT2A) serotonin receptors (Wang et al. 2009). Thus, even during chronic pain states, when the expression of these two serotonin receptor types is increased (Zhang et al. 2001, 2002), serotonergic input onto these cells is likely to lead to their increased activity, and consequently to a net decrease in spinal excitation (Wang et al. 2009).

RyRs RyRs, ER calcium release channels of which the endogenous agonist is calcium itself, can amplify calcium rises originating either intracellularly or at the plasma membrane. RyRs are likely to play a role in the transduction of pain signals by amplifying presynaptic calcium rises leading to neurotransmitter release (Huang et al. 2008; Ouyang et al. 2005). Recently, RyR activity was shown to participate in pain chronicity by stimulating the calcium-/CaM-dependent protein kinase CaMKII in nociceptive neurons (Ferrari et al. 2013). The RyR-dependent amplification of calcium signals both pre- and postsynaptically has also been proposed as an indispensible trigger for the induction LTP at C-fiber synapses in the spinal cord dorsal horn (Cheng et al. 2010; Lu et al. 2012; Naka et al. 2013). RyR-mediated internal calcium

Cell Tissue Res (2014) 357:407–426

release is implicated, as well, in pain processing along the thalamocortical pain pathway (Cheong et al. 2011), and in the mechanisms involved in central analgesia (Galeotti et al. 2005, 2008). Importantly, SERCA function and ER calcium content are reduced in sensory neurons following spinal nerve ligation and in animal models of diabetic neuropathy (Duncan et al. 2013; Gemes et al. 2009; Kruglikov et al. 2004; Rigaud et al. 2009; Zherebitskaya et al. 2012). While these findings are consistent with a role for dysregulated internal calcium release in pain sensitization, they also point toward calcium dysregulation-mediated ER stress as a proximal mechanism for neurodegeneration and the development of pain sensitivity in disorders involving peripheral nerve damage (Austin 2009; Fernyhough and Calcutt 2010).

PMCAs PMCAs represent the major high-affinity calcium extrusion mechanism in neurons, and have been identified as one of the primary regulators of presynaptic calcium at sensory neuron synapses in the spinal cord (Shutov et al. 2013). PMCA activity in primary sensory neurons is subject to regulation by rises in intracellular calcium, prolonged action potential trains, and TRPV activation, all of which accelerate PMCAmediated calcium efflux (Pottorf and Thayer 2002). The mechanisms underlying PMCA up-regulation, which might serve to limit the amplitude and duration of calcium rises in hyperactive neurons, are controlled by calcium, and mediated by CaM and PKC (Werth et al. 1996). In light of the consequences of heightened synaptic activation for the transmission of painful stimuli and for transitions to sensitized pain states, it is not inconsequential that up-regulated PMCA activity in sensory neurons promotes hyperexcitability, possibly by limiting the calcium-dependent activation of potassium and chloride channels (Gemes et al. 2012). Interestingly, PMCA expression in second order neurons within the spinal cord dorsal horn may be down-regulated in chronic pain states (Tachibana et al. 2004), and loss of PMCA expression in spinal cord neurons is linked to neuronal damage and cell loss (Kurnellas et al. 2005).

NCXs NCXs are expressed on the plasma membrane of neurons throughout the central nervous system and in cell bodies and free nerve endings of sensory neurons (Canitano et al. 2002; Persson et al. 2010). Interestingly, NCXs have two possible modes of operation depending on the electrochemical gradient of the substrate ions. Under basal conditions, they catalyze the uphill transport calcium ions out of the cell and import sodium

415

ions down their concentration gradient (forward mode). In contrast, when the plasma membrane is depolarized or when intracellular sodium concentrations are high, NCX operates in reverse mode, mediating the export of sodium and the import of calcium (Blaustein and Lederer 1999; Brini and Carafoli 2011). This feature makes the NCX a potentially very interesting subject for study in the context of the increased neuronal excitability associated with pain sensitization in general, and the sustained depolarization associated with nerve damage in particular. More specifically, although NCX typically operates in forward mode, the reverse mode of NCX operation in sensory neurons has been associated with neuropathic pain resulting from peripheral nerve injury (Kuroda et al. 2013; Muthuraman et al. 2008). Interestingly, NCXs have been localized not only at the plasma membrane of neurons but also in the inner nuclear envelope, where they may play a role in regulating nuclear and ER-associated calcium signaling (Ledeen and Wu 2007; Wu et al. 2009). The degree to which NCX and nuclear NCX contribute to pain sensitivity in neurons along pain pathways, and whether their expression and activity are altered in chronic pain states, remains to be investigated.

Mitochondrial MCU and NCLX The uptake and release of calcium ions by mitochondria, which is controlled by the activity of the MCU and NCLX proteins, respectively, represents an interesting mechanism for regulating the amplitude and duration of neuronal calcium rises and their downstream consequences (Drago et al. 2011; Shishkin et al. 2002). Indeed, calcium uptake and release by mitochondria has been proposed to control the duration of neurotransmitter release from sensory fibers in the spinal cord, where it has been identified as one of the primary regulators of presynaptic calcium transients, and to play an essential role in spinal LTP (Dedov and Roufogalis 2000; Kim et al. 2011; Medvedeva et al. 2008; Shutov et al. 2013). More specifically, the mitochondrial calcium store in capsaicin-sensitive DRG neurons was found to be functionally coupled specifically to calcium entry via TRPVs, and to facilitate the generation of long-lasting elevations in cytosolic calcium (Dedov and Roufogalis 2000; Medvedeva et al. 2008). These mitochondrially controlled, prolonged calcium rises were found, in turn, to enhance neuronal firing and neurotransmitter release (Medvedeva et al. 2008). Mitochondria were also found to buffer calcium rises resulting from DRG neuronal firing, and—together with PMCAs—to comprise one of the major presynaptic calcium clearance mechanisms in DRG/ spinal cord synapses (Shutov et al. 2013). Together, these and other findings suggest that mitochondrial calcium uptake and release control nociceptive neurotransmission in the periphery by influencing the excitability and activity of at least a

416

subset of primary sensory neurons. Mitochondria are also implicated, however, in central nociceptive sensitization. In particular, the uptake by mitochondria of calcium entering the cell via NMDARs and their subsequent generation of reactive oxygen species was found to play a key role in the signaling pathways involved in spinal LTP, inflammatory hyperalgesia, and the altered GABAergic neurotransmission associated with neuropathic pain (Kim et al. 2011; Yowtak et al. 2011). Consistent with these observations, altered mitochondrial calcium homeostasis has been implicated in the pathogenesis of chronic pain, and is linked in particular to pain sensitivity in diabetic neuropathy (Fernyhough and Calcutt 2010; Hernandez-Beltran et al. 2013; Verkhratsky and Fernyhough 2008).

CaBPs Calbindin D28k is a cytosolic EF-hand-containing CaBP that is expressed throughout the central nervous system. Although historically considered to serve as a marker for inhibitory neurons (DeFelipe 1997), calbindin D28k is also expressed by glutamatergic neurons, and this expression can be altered by excessive neuronal activity (Carter et al. 2008; Magloczky et al. 1997). Calbindin D28k is expressed at all levels of the pain axis, including the amygdala, thalamus, and spinal cord dorsal horn, and in primary sensory neurons (Craig et al. 2002; Fournet et al. 1986; Goncalves et al. 2008; Ichikawa and Sugimoto 2002; Li et al. 2005; Rausell et al. 1992). Calbindin D28k immunoreactivity has also been observed in a small population of enkephalin-positive neurons in the trigeminal subnucleus caudalis, a brainstem nucleus associated with nociception at the level of the head and thought to be the functional equivalent of the spinal cord dorsal horn (Huang et al. 2011). Consistent with a role for calbindin D28k in nociception, knockout animals were shown to exhibit deficient GABAergic neurotransmission in this nucleus and reduced nociceptive responsiveness following cutaneous injection of formalin, a substance that caused inflammatory sensitivity and increased neuronal firing in control animals (Egea et al. 2012). Moreover, increased numbers of calbindin D28kpositive nerve fibers have been linked to inflammatory sensitivity (Barcena de Arellano et al. 2013), and the presence of a substantial number of newborn calbindin D28k-positive neurons have been observed in the amygdala in an animal model of neuropathic pain (Goncalves et al. 2008; Neugebauer et al. 2004). Calbindin D28k is of course not the only cytosolic CaBP likely to play a role in nociceptive plasticity. Two of the most commonly expressed CaBPs, parvalbumin and calretinin, but also secratagogin and neuronal calcium binding proteins 1 and 2, are currently under investigation as potential players in the mechanisms underlying chronic pain (Barcena de Arellano et al. 2013; Cao et al. 2011; Hughes et al. 2012;

Cell Tissue Res (2014) 357:407–426

Pecze et al. 2013; Shi et al. 2012; Zacharova and Palecek 2009; Zhang et al. 2014).

Consequences of calcium signals Intracellular calcium rises in neurons regulate a broad spectrum of cellular processes, membrane proteins, and signaling cascades. Changes in the intracellular concentration of calcium ions can influence neuronal properties by modulating ion channel activity or localization (for instance via Homer1; see above); by regulating neurotransmitter release (for instance via the altered expression and activity of VDCCs; see above); by altering the activity of calcium-dependent enzymes, kinases, and phosphatases (Fields et al. 2005; Ji et al. 2007); and by inducing or down-regulating the expression of a wide variety of gene targets (Bading 2013; Flavell and Greenberg 2008; West et al. 2001). In light of the inhomogeneous subcellular distribution not only of calcium sources and sinks but also of calcium effectors, it is perhaps unsurprising that the physiological impact of a given type of calcium signal, or of its modification in a pathological pain state, clearly depends upon where in the cell this calcium signal occurs and from which source it derives. For instance, while a global increase in calcium influx through VDCCs in primary sensory neurons can limit neuronal excitability and thereby neurotransmitter release, a selective increase in calcium influx specifically through presynaptically localized VDCCs is likely to have the opposite effect. Thus, although primary sensory neurons express greater numbers of VDCCs in neuropathic pain models, the targeting of these channels through the α2δ subunit to presynaptic sites results in an increase, rather than a decrease, in postsynaptic excitation (Dolphin 2013; see above). Moreover, although it is by now well accepted that the development of pathological pain phenotypes requires modifications to gene expression, the subcellular localization and source of the instigating calcium signals involved determine which signaling cascades and transcriptional activators and repressors become activated, and thereby the specific pattern of genes that are affected (Bading 2013; Hagenston and Bading 2011). One particularly interesting type of neuronal calcium rise in this context is that which occurs within the nucleus (Bading 2013; Simonetti et al. 2013).

Nuclear calcium signaling Calcium rises in the nucleus can influence the transcription of specific transcription factor targets, for instance via the stimulation of calcium/CaM-dependent kinase IV (CaMKIV) and the activation of CREB and CREB binding protein (CBP), or via the activation and release of the transcriptional repressor downstream regulatory element antagonist modulator

Cell Tissue Res (2014) 357:407–426

(DREAM) from transcription factor binding sites. Nuclear calcium can also regulate gene transcription on a global level via the modulation of DNA methylation or histone acetylation, for example by altering the activity of DNA methyltransferases or by promoting the nuclear export of histone deacetylases (HDACs) (Bading 2013; Hagenston and Bading 2011; Oliveira et al. 2012; Schlumm et al. 2013; Fig. 2). Many of the targets of nuclear calcium signals are implicated in central sensitization. The transcription factor CREB, for instance, is well accepted as an important player in long-lasting spinal cord plasticity (Kuner 2010). DREAM, originally identified as a negative regulator of the Pdyn gene, which encodes the analgesic opioid polypeptide prodynorphin, has been proposed as a mediator of gene transcription and spinal cord plasticity (Cheng and Penninger 2004). Indeed, DREAM knockout animals exhibit elevated dynorphin levels and markedly reduced responses in multiple pain models (Cheng et al. 2002). Some more recent reports identifying a long-lasting up-regulation of DREAM outside the nuclear compartment of spinal cord neurons in inflammatory pain suggest, however, that the modulatory influences of DREAM on pain may be additionally attributable to mechanisms apart from the regulation of gene transcription (Long et al. 2011; Zhang et al. 2007). A particularly intriguing function for calcium signaling in general, and for nuclear calcium signaling in particular, is the epigenetic regulation of gene transcription by histone acetylation and DNA methylation (Denk and McMahon 2012). Whereas histone acetylation leads to chromatin decondensation, the deacetylation of histones by HDACs results in chromatin condensation and diminishes transcription factor accessibility. Class IIa HDACs were recently identified as targets of nuclear calcium rises. In particular, antagonism of nuclear calcium signaling in primary hippocampal neurons was shown to inhibit the nucleocytoplasmic shuttling specifically of class IIa (but not class I or class IV) HDACs. Consistent with their known involvement in transcriptional regulation, over-expression or shRNA-mediated knockdown of class IIa HDACs was also found to alter the expression of a number of known nuclear calcium regulated genes, including two that encode proteins with a proven role in central sensitization, C1qc and Cox2 (Schlumm et al. 2013; Simonetti et al. 2013). Interestingly, HDACs—including members of class II —are up-regulated in the spinal cord following the induction of inflammatory pain and down-regulated in a model of neuropathic pain (Bai et al. 2010; Tochiki et al. 2012). Moreover, their pharmacological inhibition was shown to interfere with the induction and maintenance of thermal hyperalgesia and to restore GABAergic synaptic function in the spinal cord (Bai et al. 2010; Zhang et al. 2011). Importantly, our recent paper provides the first direct evidence that nuclear calcium signals in excitatory spinal cord dorsal horn neurons play an integral role in the mechanisms

417

underlying central sensitization (Simonetti et al. 2013). More specifically, we observed that high-threshold, nociceptive-like stimulation (Torsney and MacDermott 2006) of sensory nerve fibers triggers nuclear calcium rises in spinal cord dorsal horn neurons (Fig. 2). Further, we found that the amplitude and duration of evoked nuclear calcium transients depends both on the frequency of the synaptic stimulation and on the putative number of recruited afferent axons (Simonetti et al. 2013). Thus, the increased firing rates and neurotransmitter release probability characteristic of nociceptive afferents in pathological pain states may be ideally suited to trigger nuclear calcium-dependent transcriptional responses involved in pain plasticity (Fields et al. 2005). Consistent with this idea is our finding that antagonism of nuclear calcium signaling using an exogenously expressed buffer for nuclear calcium/CaM, CaMBP4, effectively abrogates the activation of CREB, interferes with the up- or down-regulated expression of a wide range of pain-associated target genes, and disrupts the development of mechanical and thermal hypersensitivity in an animal model of chronic inflammatory pain (Simonetti et al. 2013). As outlined above, nuclear calcium signaling in spinal cord neurons is likely to regulate gene transcription and pain plasticity via multiple mechanisms (Cheng and Penninger 2002; Flavell and Greenberg 2008). Indeed, we also observed that the antagonism of CaMKIV via the overexpression of a kinase-dead CaMKIV mutant—unlike the antagonism of nuclear calcium signaling as a whole —disturbed the development of mechanical hypersensitivity, but not of thermal hypersensitivity, indicating that this pathway is differentially engaged in discrete pain modalities. Moreover, we found that only ∼30 % of genes whose expression is influenced by nuclear calcium signals in inflammatory pain are putative CREB targets (Simonetti et al. 2013). One of the primary candidate sources for nuclear calcium rises is the L-type VDCC (Hagenston and Bading 2011). Using calcium indicators expressed in the nucleus and replay of synaptic activity-evoked action potential trains, Bengtson and colleagues recently showed that L-type VDCCs in hippocampal neurons mediate at least 70 % of the amplitude of nuclear calcium rises evoked by bursts of synaptic activity (Bengtson et al. 2013). Although there is relatively little evidence implicating L-type VDCCs in the chronic pain plasticity, at least two recent studies have demonstrated that the expression of L-type VDCCs is altered in a spinal nerve ligation model of neuropathic pain (Favereaux et al. 2011; Fossat et al. 2010). In particular, these studies show that all three L-type VDCC subunits are up-regulated in chronic pain states, and that both the prevention of this up-regulation and the specific knockdown of L-type channels relieves pain hypersensitivity. They additionally observed that increased L-type VDCC expression on its own can trigger pain hypersensitivity (Favereaux et al. 2011; Fossat et al. 2010).

418

Cell Tissue Res (2014) 357:407–426

Fig. 2 Nuclear calcium signaling in pain. a The nociceptive-like electrical stimulation of C-fiber afferents evokes nuclear calcium rises in excitatory neurons of the spinal cord dorsal horn. Left schematic of the spinal cord showing the imaged area and a wide-field fluorescence image of an acute spinal cord slice expressing the FRET-based nuclear calcium indicator TnXXL.NLS. Right fluorescence traces showing nuclear calcium rises triggered by suction electrode stimulation (100 Hz, 1 s, 0.8–2.0 mA) of the attached dorsal root (adapted with permission from Simonetti et al. 2013). Bar 50 μm. b Calcium entering the nucleus of neurons can regulate gene transcription by modifying the activity of a range of target proteins and transcription factors (TFs). Consequences of nuclear calcium signaling include transcriptional disinhibition via the transcriptional

repressor DREAM, the dissociation of methyl CpG binding proteins like MeCP2 from methylated DNA, the nucleocytoplasmic export of class IIa HDACs, and the activation of CaMKIV, which in turn can phosphorylate and activate numerous targets including CREB and its cofactor CBP. Examples of genes that are up-regulated and down-regulated by nuclear calcium signaling in excitatory neurons of the spinal cord dorsal horn in inflammatory pain include CaMKIIα, Grip1, Cox2, Pdyn, and Cacnα2δ3; and Timp1, C1qc, Ctsk, C3, and Ccl3, respectively. The nuclear calcium-dependent down-regulation of C1qc disinhibits spine growth and promotes the expression of an inflammatory pain phenotype (Simonetti et al. 2013)

Importantly, using antisense and siRNA strategies, Fossat and colleagues also established a clear link between the stimulation of L-type VDCCs, the phosphorylation of CREB, and the

enhanced transcription of Cox2 that is observed following nerve ligation (Fossat et al. 2010). On the basis of these findings, it seems possible that L-type VDCC-mediated

Cell Tissue Res (2014) 357:407–426

calcium rises could represent a necessary component of excitation–transcription coupling at the spinal level.

Structural plasticity A key feature of peripheral and central sensitization is an increase in the strength of excitatory synaptic transmission at central synapses, and one of the primary mechanisms underlying this increase is a change in the density and size of synaptic spines (Kuner 2010). Rises in the intracellular concentration of calcium ions can serve as a proximal signal for such structural plasticity (Penzes et al. 2008; Simonetti et al. 2013). Several key transducers of calcium signaling that mediate pain hypersensitivity, including AMPARs and NMDARs as well as CaMKII and a number of other calcium-dependent kinases, have been shown to mediate spine stabilization and turnover, at least in brain circuits (Saneyoshi et al. 2010). In previous studies, neuropathic pain resulting from spinal cord injury was associated with increased de novo formation and elaboration of dendritic spines in spinal laminae IV and V (Tan et al. 2008). More recently, it was shown that peripheral inflammation can also lead to dendritic spine remodeling, and that this process is both dependent on calcium signaling to the nucleus and mediated by the repression of the complement cascade molecule C1q, which is known to trigger synapse elimination in mature and developing neurons (Chu et al. 2010; Ma et al. 2013; Schafer and Stevens 2010; Simonetti et al. 2013; Fig. 2). It is well established that cytoskeletal rearrangement is necessary for rapid dendritic spine plasticity, and that cytoskeletal modifications are linked to the enhanced insertion of calcium-permeable AMPARs into synapses during LTP (Fortin et al. 2010). One hypothesis is that an interaction between the AMPAR GluA2 subunit and Rac1, a member of the Rho family of GTPases, may serve to coordinate both the functional and structural plastic changes that accompany long-lasting plasticity (Asrar and Jia 2013). Several works have identified the Rho/Rac family of GTPases as key molecules involved in synaptic structural plasticity (Martinez and Tejada-Simon 2011; Schwechter et al. 2013). Indeed, these GTPases transduce signals coming from extracellular stimuli such as glutamate to the actin cytoskeleton and induce the clustering of AMPARs during spinogenesis (Schwechter and Tolias 2013; Wiens et al. 2005). Importantly, activation of Rho/Rac GTPases requires strong synaptic input, NMDAR activation, and calcium influx, and results in increased synaptic strength and synaptic spine size (Schwechter and Tolias 2013). Further, activation of Rac1 following spinal cord injury mediates changes in spine morphology and density in the spinal dorsal horn, and leads to injury-induced hyperalgesia (Tan 2008). Moreover, as discussed above, GluA2 AMPAR subunits are internalized in animal models of persistent inflammatory and neuropathic

419

pain (Chen et al. 2013; Larsson and Broman 2008; Park et al. 2009), and the absence of GluA2 prevents the synapse loss that is otherwise observed following the mGluR-dependent long-term depression of synaptic strength (Nosyreva and Huber 2005). Therefore, the combined calcium-dependent up-regulation of Rac1 and internalization of GluA2 are likely to be permissive for structural plasticity in the spinal cord dorsal horn.

Outlook While it is well accepted that the amplitude, kinetics, and distribution of a given calcium rise are important for determining what its consequences might be (Hagenston and Bading 2011; Hardingham and Bading 2010; Raymond and Redman 2006), surprisingly little is known about the characteristics of synaptic activity-triggered calcium signals in distinct subcellular compartments, their contributions to cellular and functional plasticity, and whether and how they are altered in acute and chronic pain states. Moreover, although research into the mechanisms of chronic pain sensitization has expanded in recent years and includes investigations not only of excitatory neurons but also of astrocytes and microglia (Bardoni et al. 2010; Beggs and Salter 2010; Gao and Ji 2010; McMahon and Malcangio 2009), exceptionally little is known about their activation patterns in response to painful stimuli or about the calcium signals involved in these processes. Further, although intracellular calcium rises have been shown to be important for intercellular communication in activated astrocytes and microglia (Bardoni et al. 2010; Beggs and Salter 2010; Gao and Ji 2010; McMahon and Malcangio 2009), whether calcium signaling mediates transitions of glial cells from resting to activated states has not yet been examined. Finally, the characteristics of calcium signals in distinct types of neurons, both excitatory and inhibitory, throughout the entire pain axis have not been widely studied. A number of recent improvements in the properties of genetically encoded calcium indicators, however, should make possible the direct examination and characterization of cell-type- and subcellular compartment-specific calcium signals in chronic pain states (Akerboom et al. 2013; Ohkura et al. 2012). When employed in concert with pharmacological tools to modify specific ion channels or pathways, and/or with molecular tools that enable the knockdown or over-expression of select target proteins (Bengtson et al. 2010; Qiu et al. 2013; Simonetti et al. 2013), the use of such indicators offers a powerful means to explore the characteristics and consequences of global and subcellularly localized neuronal and glial calcium signals in pathological pain states.

420

References Akerboom J, Carreras Calderon N, Tian L, Wabnig S, Prigge M, Tolo J, Gordus A, Orger MB, Severi KE, Macklin JJ, Patel R, Pulver SR, Wardill TJ, Fischer E, Schuler C, Chen TW, Sarkisyan KS, Marvin JS, Bargmann CI, Kim DS, Kugler S, Lagnado L, Hegemann P, Gottschalk A, Schreiter ER, Looger LL (2013) Genetically encoded calcium indicators for multi-color neural activity imaging and combination with optogenetics. Front Mol Neurosci 6:2 Amir R, Devor M (1997) Spike-evoked suppression and burst patterning in dorsal root ganglion neurons of the rat. J Physiol 501(Pt 1):183– 196 Ango F, Prezeau L, Muller T, Tu JC, Xiao B, Worley PF, Pin JP, Bockaert J, Fagni L (2001) Agonist-independent activation of metabotropic glutamate receptors by the intracellular protein homer. Nature 411: 962–965 Asrar S, Jia Z (2013) Molecular mechanisms coordinating functional and morphological plasticity at the synapse: role of GluA2/N-cadherin interaction-mediated actin signaling in mGluR-dependent LTD. Cell Signal 25:397–402 Austin RC (2009) The unfolded protein response in health and disease. Antioxid Redox Signal 11:2279–2287 Bading H (2013) Nuclear calcium signalling in the regulation of brain function. Nat Rev Neurosci 14:593–608 Bahia PK, Suzuki R, Benton DC, Jowett AJ, Chen MX, Trezise DJ, Dickenson AH, Moss GW (2005) A functional role for smallconductance calcium-activated potassium channels in sensory pathways including nociceptive processes. J Neurosci 25:3489–3498 Bai G, Wei D, Zou S, Ren K, Dubner R (2010) Inhibition of class II histone deacetylases in the spinal cord attenuates inflammatory hyperalgesia. Mol Pain 6:51 Barcena de Arellano ML, Munch S, Arnold J, Helbig S, Schneider A, Mechsner S (2013) Calcium-binding protein expression in peritoneal endometriosis-associated nerve fibres. Eur J Pain 17:1425– 1437 Bardoni R, Ghirri A, Zonta M, Betelli C, Vitale G, Ruggieri V, Sandrini M, Carmignoto G (2010) Glutamate-mediated astrocyte-to-neuron signalling in the rat dorsal horn. J Physiol 588:831–846 Basbaum AI, Bautista DM, Scherrer G, Julius D (2009) Cellular and molecular mechanisms of pain. Cell 139:267–284 Bauer CS, Nieto-Rostro M, Rahman W, Tran-Van-Minh A, Ferron L, Douglas L, Kadurin I, Sri Ranjan Y, Fernandez- Alacid L, Millar NS, Dickenson AH, Lujan R, Dolphin AC (2009) The increased trafficking of the calcium channel subunit alpha2delta-1 to presynaptic terminals in neuropathic pain is inhibited by the alpha2delta ligand pregabalin. J Neurosci 29:4076–4088 Beggs S, Salter MW (2010) Microglia-neuronal signalling in neuropathic pain hypersensitivity 2.0. Curr Opin Neurobiol 20:474–480 Begon S, Pickering G, Eschalier A, Dubray C (2000) Magnesium and MK-801 have a similar effect in two experimental models of neuropathic pain. Brain Res 887:436–439 Bengtson CP, Freitag HE, Weislogel JM, Bading H (2010) Nuclear calcium sensors reveal that repetition of trains of synaptic stimuli boosts nuclear calcium signaling in CA1 pyramidal neurons. Biophys J 99:4066–4077 Bengtson CP, Kaiser M, Obermayer J, Bading H (2013) Calcium responses to synaptically activated bursts of action potentials and their synapse-independent replay in cultured networks of hippocampal neurons. Biochim Biophys Acta (in press) Bernier LP, Ase AR, Seguela P (2013) Post-translational regulation of P2X receptor channels: modulation by phospholipids. Front Cell Neurosci 7:226 Berridge MJ (1998) Neuronal calcium signaling. Neuron 21:13–26 Blaustein MP, Lederer WJ (1999) Sodium/calcium exchange: its physiological implications. Physiol Rev 79:763–854

Cell Tissue Res (2014) 357:407–426 Boroujerdi A, Kim HK, Lyu YS, Kim DS, Figueroa KW, Chung JM, Luo ZD (2008) Injury discharges regulate calcium channel alpha-2-delta1 subunit upregulation in the dorsal horn that contributes to initiation of neuropathic pain. Pain 139:358–366 Boroujerdi A, Zeng J, Sharp K, Kim D, Steward O, Luo ZD (2011) Calcium channel alpha-2-delta-1 protein upregulation in dorsal spinal cord mediates spinal cord injury-induced neuropathic pain states. Pain 152:649–655 Borvendeg SJ, Gerevich Z, Gillen C, Illes P (2003) P2Y receptormediated inhibition of voltage-dependent Ca2+ channels in rat dorsal root ganglion neurons. Synapse 47:159–161 Bourinet E, Alloui A, Monteil A, Barrere C, Couette B, Poirot O, Pages A, McRory J, Snutch TP, Eschalier A, Nargeot J (2005) Silencing of the Cav3.2 T-type calcium channel gene in sensory neurons demonstrates its major role in nociception. EMBO J 24:315–324 Bourinet E, Altier C, Hildebrand ME, Trang T, Salter MW, Zamponi GW (2014) Calcium-permeable ion channels in pain signaling. Physiol Rev 94:81–140 Brini M, Carafoli E (2011) The plasma membrane Ca (2)+ATPase and the plasma membrane sodium calcium exchanger cooperate in the regulation of cell calcium. Cold Spring Harb Perspect Biol 3 Campbell JN, Meyer RA (2006) Mechanisms of neuropathic pain. Neuron 52:77–92 Canitano A, Papa M, Boscia F, Castaldo P, Sellitti S, Taglialatela M, Annunziato L (2002) Brain distribution of the Na+/Ca2+ exchangerencoding genes NCX1, NCX2, and NCX3 and their related proteins in the central nervous system. Ann NY Acad Sci 976:394–404 Cao MH, Ji FT, Liu L, Li F (2011) Expression changes of parvalbumin and microtubule-associated protein 2 induced by chronic constriction injury in rat dorsal root ganglia. Chin Med J Engl 124:2184– 2190 Carter DS, Harrison AJ, Falenski KW, Blair RE, DeLorenzo RJ (2008) Long-term decrease in calbindin-D28K expression in the hippocampus of epileptic rats following pilocarpine-induced status epilepticus. Epilepsy Res 79:213–223 Chen SR, Zhou HY, Byun HS, Pan HL (2013) Nerve injury increases GluA2-lacking AMPA receptor prevalence in spinal cords: functional significance and signaling mechanisms. J Pharmacol Exp Ther 347:765–772 Cheng HY, Penninger JM (2002) Transcriptional mechanisms underlying neuropathic pain: DREAM, transcription factors and future pain management? Expert Rev Neurother 2:677–689 Cheng HY, Penninger JM (2004) DREAMing about arthritic pain. Ann Rheum Dis 63 Suppl 2:ii72-ii75 Cheng HY, Pitcher GM, Laviolette SR, Whishaw IQ, Tong KI, Kockeritz LK, Wada T, Joza NA, Crackower M, Goncalves J, Sarosi I, Woodgett JR, Oliveira-dos-Santos AJ, Ikura M, van der Kooy D, Salter MW, Penninger JM (2002) DREAM is a critical transcriptional repressor for pain modulation. Cell 108:31–43 Cheng LZ, Lu N, Zhang YQ, Zhao ZQ (2010) Ryanodine receptors contribute to the induction of nociceptive input-evoked long-term potentiation in the rat spinal cord slice. Mol Pain 6:1 Cheong E, Shin HS (2013) T-type Ca2+ channels in normal and abnormal brain functions. Physiol Rev 93:961–992 Cheong E, Lee S, Choi BJ, Sun M, Lee CJ, Shin HS (2008) Tuning thalamic firing modes via simultaneous modulation of T- and L-type Ca2+ channels controls pain sensory gating in the thalamus. J Neurosci 28:13331–13340 Cheong E, Kim C, Choi BJ, Sun M, Shin HS (2011) Thalamic ryanodine receptors are involved in controlling the tonic firing of thalamocortical neurons and inflammatory pain signal processing. J Neurosci 31:1213–1218 Chizh BA, Headley PM (2005) NMDA antagonists and neuropathic pain–multiple drug targets and multiple uses. Curr Pharm Des 11: 2977–2994

Cell Tissue Res (2014) 357:407–426 Choi S, Na HS, Kim J, Lee J, Lee S, Kim D, Park J, Chen CC, Campbell KP, Shin HS (2007) Attenuated pain responses in mice lacking Ca (V)3.2 T-type channels. Genes Brain Behav 6:425–431 Chu Y, Jin X, Parada I, Pesic A, Stevens B, Barres B, Prince DA (2010) Enhanced synaptic connectivity and epilepsy in C1q knockout mice. Proc Natl Acad Sci USA 107:7975–7980 Chun YH, Frank D, Lee JS, Zhang Y, Auh QS, Ro JY (2008) Peripheral AMPA receptors contribute to muscle nociception and c-fos activation. Neurosci Res 62:97–104 Cizkova D, Marsala J, Lukacova N, Marsala M, Jergova S, Orendacova J, Yaksh TL (2002) Localization of N-type Ca2+ channels in the rat spinal cord following chronic constrictive nerve injury. Exp Brain Res 147:456–463 Craig AD, Zhang ET, Blomqvist A (2002) Association of spinothalamic lamina I neurons and their ascending axons with calbindinimmunoreactivity in monkey and human. Pain 97:105–115 Cruz CD, Cruz F (2007) The ERK 1 and 2 pathway in the nervous system: from basic aspects to possible clinical applications in pain and visceral dysfunction. Curr Neuropharmacol 5:244–252 Dang K, Bielfeldt K, Lamb K, Gebhart GF (2005) Gastric ulcers evoke hyperexcitability and enhance P2X receptor function in rat gastric sensory neurons. J Neurophysiol 93:3112–3119 D’Arco M, Giniatullin R, Simonetti M, Fabbro A, Nair A, Nistri A, Fabbretti E (2007) Neutralization of nerve growth factor induces plasticity of ATP-sensitive P2X3 receptors of nociceptive trigeminal ganglion neurons. J Neurosci 27:8190–8201 Davies A, Hendrich J, Van Minh AT, Wratten J, Douglas L, Dolphin AC (2007) Functional biology of the alpha (2) delta subunits of voltage-gated calcium channels. Trends Pharmacol Sci 28: 220–228 Dedov VN, Roufogalis BD (2000) Mitochondrial calcium accumulation following activation of vanilloid (VR1) receptors by capsaicin in dorsal root ganglion neurons. Neuroscience 95:183–188 DeFelipe J (1997) Types of neurons, synaptic connections and chemical characteristics of cells immunoreactive for calbindin-D28K, parvalbumin and calretinin in the neocortex. J Chem Neuroanat 14:1–19 Denk F, McMahon SB (2012) Chronic pain: emerging evidence for the involvement of epigenetics. Neuron 73:435–444 Dolphin AC (2013) The alpha2delta subunits of voltage-gated calcium channels. Biochim Biophys Acta 1828:1541–1549 Donnelly-Roberts D, McGaraughty S, Shieh CC, Honore P, Jarvis MF (2008) Painful purinergic receptors. J Pharmacol Exp Ther 324: 409–415 Drago I, Pizzo P, Pozzan T (2011) After half a century mitochondrial calcium in- and efflux machineries reveal themselves. EMBO J 30: 4119–4125 Duncan C, Mueller S, Simon E, Renger JJ, Uebele VN, Hogan QH, Wu HE (2013) Painful nerve injury decreases sarco- endoplasmic reticulum Ca (2) (+)-ATPase activity in axotomized sensory neurons. Neuroscience 231:247–257 Egea J, Malmierca E, Rosa AO, del Barrio L, Negredo P, Nunez A, Lopez MG (2012) Participation of calbindin-D28K in nociception: results from calbindin-D28K knockout mice. Pflugers Arch 463:449–458 Fabbretti E (2013) ATP P2X3 receptors and neuronal sensitization. Front Cell Neurosci 7:236 Favereaux A, Thoumine O, Bouali-Benazzouz R, Roques V, Papon MA, Salam SA, Drutel G, Leger C, Calas A, Nagy F, Landry M (2011) Bidirectional integrative regulation of Cav1.2 calcium channel by microRNA miR-103: role in pain. EMBO J 30: 3830–3841 Felsby S, Nielsen J, Arendt-Nielsen L, Jensen TS (1996) NMDA receptor blockade in chronic neuropathic pain: a comparison of ketamine and magnesium chloride. Pain 64:283–291 Fernyhough P, Calcutt NA (2010) Abnormal calcium homeostasis in peripheral neuropathies. Cell Calcium 47:130–139

421 Ferrari LF, Bogen O, Levine JD (2013) Role of nociceptor alphaCaMKII in transition from acute to chronic pain (hyperalgesic priming) in male and female rats. J Neurosci 33:11002–11011 Fields RD, Lee PR, Cohen JE (2005) Temporal integration of intracellular Ca2+ signaling networks in regulating gene expression by action potentials. Cell Calcium 37:433–442 Fisher K, Coderre TJ, Hagen NA (2000) Targeting the N-methyl-Daspartate receptor for chronic pain management. preclinical animal studies, recent clinical experience and future research directions. J Pain Symptom Manag 20:358–373 Flake NM, Bonebreak DB, Gold MS (2005) Estrogen and inflammation increase the excitability of rat temporomandibular joint afferent neurons. J Neurophysiol 93:1585–1597 Flavell SW, Greenberg ME (2008) Signaling mechanisms linking neuronal activity to gene expression and plasticity of the nervous system. Annu Rev Neurosci 31:563–590 Fortin DA, Davare MA, Srivastava T, Brady JD, Nygaard S, Derkach VA, Soderling TR (2010) Long-term potentiationdependent spine enlargement requires synaptic Ca2 + −permeable AMPA receptors recruited by CaM-kinase I. J Neurosci 30:11565–11575 Fossat P, Dobremez E, Bouali-Benazzouz R, Favereaux A, Bertrand SS, Kilk K, Leger C, Cazalets JR, Langel U, Landry M, Nagy F (2010) Knockdown of L calcium channel subtypes: differential effects in neuropathic pain. J Neurosci 30:1073–1085 Fournet N, Garcia-Segura LM, Norman AW, Orci L (1986) Selective localization of calcium-binding protein in human brainstem, cerebellum and spinal cord. Brain Res 399:310–316 Fuchs A, Lirk P, Stucky C, Abram SE, Hogan QH (2005) Painful nerve injury decreases resting cytosolic calcium concentrations in sensory neurons of rats. Anesthesiology 102:1217–1225 Fuchs A, Rigaud M, Hogan QH (2007) Painful nerve injury shortens the intracellular Ca2+ signal in axotomized sensory neurons of rats. Anesthesiology 107:106–116 Galan A, Laird JM, Cervero F (2004) In vivo recruitment by painful stimuli of AMPA receptor subunits to the plasma membrane of spinal cord neurons. Pain 112:315–323 Galeotti N, Bartolini A, Ghelardini C (2005) Ryanodine receptors are involved in muscarinic antinociception in mice. Behav Brain Res 164:165–171 Galeotti N, Quattrone A, Vivoli E, Bartolini A, Ghelardini C (2008) Type 1 and type 3 ryanodine receptors are selectively involved in muscarinic antinociception in mice: an antisense study. Neuroscience 153: 814–822 Gangadharan V, Wang R, Ulzhofer B, Luo C, Bardoni R, Bali KK, Agarwal N, Tegeder I, Hildebrandt U, Nagy GG, Todd AJ, Ghirri A, Haussler A, Sprengel R, Seeburg PH, MacDermott AB, Lewin GR, Kuner R (2011) Peripheral calcium- permeable AMPA receptors regulate chronic inflammatory pain in mice. J Clin Invest 121: 1608–1623 Gao YJ, Ji RR (2010) Targeting astrocyte signaling for chronic pain. Neurotherapeutics 7:482–493 Gemes G, Rigaud M, Weyker PD, Abram SE, Weihrauch D, Poroli M, Zoga V, Hogan QH (2009) Depletion of calcium stores in injured sensory neurons: anatomic and functional correlates. Anesthesiology 111:393–405 Gemes G, Oyster KD, Pan B, Wu HE, Bangaru ML, Tang Q, Hogan QH (2012) Painful nerve injury increases plasma membrane Ca2+ −ATPase activity in axotomized sensory neurons. Mol Pain 8:46 Gerevich Z, Muller C, Illes P (2005) Metabotropic P2Y1 receptors inhibit P2X3 receptor-channels in rat dorsal root ganglion neurons. Eur J Pharmacol 521:34–38 Gerevich Z, Zadori Z, Muller C, Wirkner K, Schroder W, Rubini P, Illes P (2007) Metabotropic P2Y receptors inhibit P2X3 receptor-channels via G protein-dependent facilitation of their desensitization. Br J Pharmacol 151:226–236

422 Giniatullin R, Nistri A, Fabbretti E (2008) Molecular mechanisms of sensitization of pain-transducing P2X3 receptors by the migraine mediators CGRP and NGF. Mol Neurobiol 37:83–90 Gnanasekaran A, Sundukova M, Hullugundi S, Birsa N, Bianchini G, Hsueh YP, Nistri A, Fabbretti E (2013) Calcium/ calmodulin-dependent serine protein kinase (CASK) is a new intracellular modulator of P2X3 receptors. J Neurochem 126: 102–112 Goncalves L, Silva R, Pinto-Ribeiro F, Pego JM, Bessa JM, Pertovaara A, Sousa N, Almeida A (2008) Neuropathic pain is associated with depressive behaviour and induces neuroplasticity in the amygdala of the rat. Exp Neurol 213:48–56 Gu JG, Albuquerque C, Lee CJ, MacDermott AB (1996) Synaptic strengthening through activation of Ca2+−permeable AMPA receptors. Nature 381:793–796 Hagenston AM, Bading H (2011) Calcium signaling in synapse-tonucleus communication. Cold Spring Harb Perspect Biol 3:a004564 Hall KE, Sima AA, Wiley JW (1995) Voltage-dependent calcium currents are enhanced in dorsal root ganglion neurones from the bio bred/worchester diabetic rat. J Physiol 486(Pt 2):313–322 Hardingham GE (2009) Coupling of the NMDA receptor to neuroprotective and neurodestructive events. Biochem Soc Trans 37:1147– 1160 Hardingham GE, Bading H (2010) Synaptic versus extrasynaptic NMDA receptor signalling: implications for neurodegenerative disorders. Nat Rev Neurosci 11:682–696 Hartmann B, Ahmadi S, Heppenstall PA, Lewin GR, Schott C, Borchardt T, Seeburg PH, Zeilhofer HU, Sprengel R, Kuner R (2004) The AMPA receptor subunits GluR-A and GluR-B reciprocally modulate spinal synaptic plasticity and inflammatory pain. Neuron 44: 637–650 Heinricher MM, Tavares I, Leith JL, Lumb BM (2009) Descending control of nociception: specificity, recruitment and plasticity. Brain Res Rev 60:214–225 Hendrich J, Van Minh AT, Heblich F, Nieto-Rostro M, Watschinger K, Striessnig J, Wratten J, Davies A, Dolphin AC (2008) Pharmacological disruption of calcium channel trafficking by the alpha2delta ligand gabapentin. Proc Natl Acad Sci USA 105:3628– 3633 Hernandez-Beltran N, Moreno CB, Gutierrez-Alvarez AM (2013) Contribution of mitochondria to pain in diabetic neuropathy. Endocrinol Nutr Organo Soc Esp Endocrinol Nutr 60:25–32 Ho YC, Cheng JK, Chiou LC (2013) Hypofunction of glutamatergic neurotransmission in the periaqueductal gray contributes to nerveinjury-induced neuropathic pain. J Neurosci 33:7825–7836 Hogan QH (2007) Role of decreased sensory neuron membrane calcium currents in the genesis of neuropathic pain. Croat Med J 48:9–21 Hogan Q, Lirk P, Poroli M, Rigaud M, Fuchs A, Fillip P, Ljubkovic M, Gemes G, Sapunar D (2008) Restoration of calcium influx corrects membrane hyperexcitability in injured rat dorsal root ganglion neurons. Anesth Analg 107:1045–1051 Hollmann M, Hartley M, Heinemann S (1991) Ca2+ permeability of KAAMPA–gated glutamate receptor channels depends on subunit composition. Science 252:851–853 Hoppa MB, Lana B, Margas W, Dolphin AC, Ryan TA (2012) alpha2delta expression sets presynaptic calcium channel abundance and release probability. Nature 486:122–125 Huang W, Wang H, Galligan JJ, Wang DH (2008) Transient receptor potential vanilloid subtype 1 channel mediated neuropeptide secretion and depressor effects: role of endoplasmic reticulum associated Ca2+ release receptors in rat dorsal root ganglion neurons. J Hypertens 26:1966–1975 Huang J, Wang W, Chen J, Ge SN, Wei YY, Wang YY, Kaneko T, Li YQ, Wu SX (2011) Neurochemical features of enkephalinergic neurons in the mouse trigeminal subnucleus caudalis. Neurochem Int 58:44–51

Cell Tissue Res (2014) 357:407–426 Hughes DI, Sikander S, Kinnon CM, Boyle KA, Watanabe M, Callister RJ, Graham BA (2012) Morphological, neurochemical and electrophysiological features of parvalbumin-expressing cells: a likely source of axo-axonic inputs in the mouse spinal dorsal horn. J Physiol 590:3927–3951 Ichikawa H, Sugimoto T (2002) Co-expression of VRL-1 and calbindin D-28k in the rat sensory ganglia. Brain Res 924:109–112 Jagodic MM, Pathirathna S, Nelson MT, Mancuso S, Joksovic PM, Rosenberg ER, Bayliss DA, Jevtovic-Todorovic V, Todorovic SM (2007) Cell-specific alterations of T-type calcium current in painful diabetic neuropathy enhance excitability of sensory neurons. J Neurosci 27:3305–3316 Jardin I, Lopez JJ, Berna-Erro A, Salido GM, Rosado JA (2013) Homer proteins in Ca (2) (+) entry. IUBMB life 65:497–504 Jarvis MF (2010) The neural-glial purinergic receptor ensemble in chronic pain states. Trends Neurosci 33:48–57 Ji RR, Kohno T, Moore KA, Woolf CJ (2003) Central sensitization and LTP: do pain and memory share similar mechanisms? Trends Neurosci 26:696–705 Ji RR, Kawasaki Y, Zhuang ZY, Wen YR, Zhang YQ (2007) Protein kinases as potential targets for the treatment of pathological pain. Handb Exp Pharmacol 359–389 Jin X, Shah S, Liu Y, Zhang H, Lees M, Fu Z, Lippiat JD, Beech DJ, Sivaprasadarao A, Baldwin SA, Zhang H, Gamper N (2013) Activation of the Cl- channel ANO1 by localized calcium signals in nociceptive sensory neurons requires coupling with the IP3 receptor. Sci Signal 6:ra73 Kaczmarek-Hajek K, Lorinczi E, Hausmann R, Nicke A (2012) Molecular and functional properties of P2X receptors– recent progress and persisting challenges. Purinergic Signal 8:375–417 Katz EJ, Gold MS (2006) Inflammatory hyperalgesia: a role for the Cfiber sensory neuron cell body? J Pain 7:170–178 Khomula EV, Viatchenko-Karpinski VY, Borisyuk AL, Duzhyy DE, Belan PV, Voitenko NV (2013) Specific functioning of Cav3.2 T-type calcium and TRPV1 channels under different types of STZ-diabetic neuropathy. Biochim Biophys Acta 1832:636–649 Khomula EV, Borisyuk AL, Viatchenko-Karpinski VY, Briede A, Belan PV, Voitenko NV (2014) Nociceptive neurons differentially express fast and slow T-type Ca (2) (+) currents in different types of diabetic neuropathy. Neural Plast 2014:938235 Kim C, Chung JM, Chung K (2003) Changes in the gene expression of six subtypes of P2X receptors in rat dorsal root ganglion after spinal nerve ligation. Neurosci Lett 337:81–84 Kim HY, Lee KY, Lu Y, Wang J, Cui L, Kim SJ, Chung JM, Chung K (2011) Mitochondrial Ca (2+) uptake is essential for synaptic plasticity in pain. J Neurosci 31:12982–12991 Kopach O, Viatchenko-Karpinski V, Atianjoh FE, Belan P, Tao Y-X, Voitenko N (2013) PKC alpha Is required for inflammationinduced trafficking of extrasynaptic AMPA receptors in tonically firing lamina II dorsal horn neurons during the maintenance of persistent inflammatory pain. J Pain 14:182–192 Kruglikov I, Gryshchenko O, Shutov L, Kostyuk E, Kostyuk P, Voitenko N (2004) Diabetes-induced abnormalities in ER calcium mobilization in primary and secondary nociceptive neurons. Pflugers Arch 448:395–401 Kuner R (2010) Central mechanisms of pathological pain. Nat Med 16: 1258–1266 Kurnellas MP, Nicot A, Shull GE, Elkabes S (2005) Plasma membrane calcium ATPase deficiency causes neuronal pathology in the spinal cord: a potential mechanism for neurodegeneration in multiple sclerosis and spinal cord injury. FASEB J 19:298–300 Kuroda H, Shibukawa Y, Soya M, Masamura A, Kasahara M, Tazaki M, Ichinohe T (2012) Expression of P2X (1) and P2X (4) receptors in rat trigeminal ganglion neurons. Neuroreport 23:752–756

Cell Tissue Res (2014) 357:407–426 Kuroda H, Sobhan U, Sato M, Tsumura M, Ichinohe T, Tazaki M, Shibukawa Y (2013) Sodium-calcium exchangers in rat trigeminal ganglion neurons. Mol Pain 9:22 Kweon HJ, Suh BC (2013) Acid-sensing ion channels (ASICs): therapeutic targets for neurological diseases and their regulation. BMB Rep 46:295–304 Larsson M, Broman J (2008) Translocation of GluR1-containing AMPA receptors to a spinal nociceptive synapse during acute noxious stimulation. J Neurosci 28:7084–7090 Latham JR, Pathirathna S, Jagodic MM, Choe WJ, Levin ME, Nelson MT, Lee WY, Krishnan K, Covey DF, Todorovic SM, Jevtovic-Todorovic V (2009) Selective T-type calcium channel blockade alleviates hyperalgesia in ob/ob mice. Diabetes 58:2656–2665 Ledeen R, Wu G (2007) GM1 in the nuclear envelope regulates nuclear calcium through association with a nuclear sodium- calcium exchanger. J Neurochem 103(Suppl 1):126–134 Leong ML, Gu M, Speltz-Paiz R, Stahura EI, Mottey N, Steer CJ, Wessendorf M (2011) Neuronal loss in the rostral ventromedial medulla in a rat model of neuropathic pain. J Neurosci 31:17028– 17039 Li CY, Song YH, Higuera ES, Luo ZD (2004) Spinal dorsal horn calcium channel alpha2delta-1 subunit upregulation contributes to peripheral nerve injury-induced tactile allodynia. J Neurosci 24:8494–8499 Li YN, Sakamoto H, Kawate T, Cheng CX, Li YC, Shimada O, Atsumi S (2005) An immunocytochemical study of calbindin-D28K in laminae I and II of the dorsal horn and spinal ganglia in the chicken with special reference to the relation to substance P-containing primary afferent neurons. Arch Histol Cytol 68:57–70 Li CY, Zhang XL, Matthews EA, Li KW, Kurwa A, Boroujerdi A, Gross J, Gold MS, Dickenson AH, Feng G, Luo ZD (2006) Calcium channel alpha2delta1 subunit mediates spinal hyperexcitability in pain modulation. Pain 125:20–34 Li KW, Yu YP, Zhou C, Kim DS, Lin B, Sharp K, Steward O, Luo ZD (2014a) Calcium channel alpha2delta1 proteins mediate trigeminal neuropathic pain states associated with aberrant excitatory synaptogenesis. J Biol Chem Li N, Lu ZY, Yu LH, Burnstock G, Deng XM, Ma B (2014b) Inhibition of G protein-coupled P2Y2 receptor induced analgesia in a rat model of trigeminal neuropathic pain. Mol Pain 10:21 Lirk P, Poroli M, Rigaud M, Fuchs A, Fillip P, Huang CY, Ljubkovic M, Sapunar D, Hogan Q (2008) Modulators of calcium influx regulate membrane excitability in rat dorsal root ganglion neurons. Anesth Analg 107:673–685 Lisman JE (1997) Bursts as a unit of neural information: making unreliable synapses reliable. Trends Neurosci 20:38–43 Liu XJ, Salter MW (2010) Glutamate receptor phosphorylation and trafficking in pain plasticity in spinal cord dorsal horn. Eur J Neurosci 32:278–289 Long I, Suppian R, Ismail Z (2011) Increases in mRNA and DREAM protein expression in the rat spinal cord after formalin induced pain. Neurochem Res 36:533–539 Lu SG, Gold MS (2008) Inflammation-induced increase in evoked calcium transients in subpopulations of rat dorsal root ganglion neurons. Neuroscience 153:279–288 Lu SG, Zhang XL, Luo ZD, Gold MS (2010) Persistent inflammation alters the density and distribution of voltage- activated calcium channels in subpopulations of rat cutaneous DRG neurons. Pain 151:633–643 Lu N, Cheng LZ, Zhang YQ, Lu BC, Li YQ, Zhao ZQ (2012) Involvement of ryanodine receptors in tetanic sciatic stimulationinduced long-term potentiation of spinal dorsal horn and persistent pain in rats. J Neurosci Res 90:1096–1104 Luo C, Seeburg PH, Sprengel R, Kuner R (2008) Activity-dependent potentiation of calcium signals in spinal sensory networks in inflammatory pain states. Pain 140:358–367

423 Ma W, Quirion R (2014) Targeting cell surface trafficking of painfacilitating receptors to treat chronic pain conditions. Expert Opin Ther Targets 18:459–472 Ma Y, Ramachandran A, Ford N, Parada I, Prince DA (2013) Remodeling of dendrites and spines in the C1q knockout model of genetic epilepsy. Epilepsia 54:1232–1239 Magloczky Z, Halasz P, Vajda J, Czirjak S, Freund TF (1997) Loss of Calbindin-D28K immunoreactivity from dentate granule cells in human temporal lobe epilepsy. Neuroscience 76:377–385 Malin SA, Molliver DC (2010) Gi- and Gq-coupled ADP (P2Y) receptors act in opposition to modulate nociceptive signaling and inflammatory pain behavior. Mol Pain 6:21 Martinez LA, Tejada-Simon MV (2011) Pharmacological inactivation of the small GTPase Rac1 impairs long-term plasticity in the mouse hippocampus. Neuropharmacology 61:305–312 Mattson MP (2012) Parkinson’s disease: don’t mess with calcium. J Clin Invest 122:1195–1198 McCallum JB, Kwok WM, Mynlieff M, Bosnjak ZJ, Hogan QH (2003) Loss of T-type calcium current in sensory neurons of rats with neuropathic pain. Anesthesiology 98:209–216 McCallum JB, Kwok WM, Sapunar D, Fuchs A, Hogan QH (2006) Painful peripheral nerve injury decreases calcium current in axotomized sensory neurons. Anesthesiology 105:160–168 McCallum JB, Wu HE, Tang Q, Kwok WM, Hogan QH (2011) Subtypespecific reduction of voltage-gated calcium current in medium-sized dorsal root ganglion neurons after painful peripheral nerve injury. Neuroscience 179:244–255 McKelvey L, Shorten GD, O’Keeffe GW (2013) Nerve growth factormediated regulation of pain signalling and proposed new intervention strategies in clinical pain management. J Neurochem 124:276– 289 McMahon SB, Malcangio M (2009) Current challenges in glia-pain biology. Neuron 64:46–54 Medvedeva YV, Kim MS, Usachev YM (2008) Mechanisms of prolonged presynaptic Ca2+ signaling and glutamate release induced by TRPV1 activation in rat sensory neurons. J Neurosci 28: 5295–5311 Messinger RB, Naik AK, Jagodic MM, Nelson MT, Lee WY, Choe WJ, Orestes P, Latham JR, Todorovic SM, Jevtovic- Todorovic V (2009) In vivo silencing of the Ca (V)3.2 T-type calcium channels in sensory neurons alleviates hyperalgesia in rats with streptozocininduced diabetic neuropathy. Pain 145:184–195 Millan MJ (1999) The induction of pain: an integrative review. Prog Neurobiol 57:1–164 Millan MJ (2002) Descending control of pain. Prog Neurobiol 66:355– 474 Miyabe T, Miletic G, Miletic V (2006) Loose ligation of the sciatic nerve in rats elicits transient up-regulation of Homer1a gene expression in the spinal dorsal horn. Neurosci Lett 398:296–299 Moore BA, Stewart TM, Hill C, Vanner SJ (2002) TNBS ileitis evokes hyperexcitability and changes in ionic membrane properties of nociceptive DRG neurons. Am J Physiol Gastrointest Liver Physiol 282:G1045–1051 Moriyama T, Iida T, Kobayashi K, Higashi T, Fukuoka T, Tsumura H, Leon C, Suzuki N, Inoue K, Gachet C, Noguchi K, Tominaga M (2003) Possible involvement of P2Y2 metabotropic receptors in ATP-induced transient receptor potential vanilloid receptor 1-mediated thermal hypersensitivity. J Neurosci 23:6058– 6062 Moseley GL, Flor H (2012) Targeting cortical representations in the treatment of chronic pain: a review. Neurorehabil Neural Repair 26:646–652 Muthuraman A, Jaggi AS, Singh N, Singh D (2008) Ameliorative effects of amiloride and pralidoxime in chronic constriction injury and vincristine induced painful neuropathy in rats. Eur J Pharmacol 587:104–111

424 Naka A, Gruber-Schoffnegger D, Sandkühler J (2013) Non-Hebbian plasticity at C-fiber synapses in rat spinal cord lamina I neurons. Pain 154:1333–1342 Naziroglu M, Dikici DM, Dursun S (2012) Role of oxidative stress and Ca (2) (+) signaling on molecular pathways of neuropathic pain in diabetes: focus on TRP channels. Neurochem Res 37:2065–2075 Nelson MT, Todorovic SM (2006) Is there a role for T-type calcium channels in peripheral and central pain sensitization? Mol Neurobiol 34:243–248 Neugebauer V, Li W, Bird GC, Han JS (2004) The amygdala and persistent pain. Neuroscientist 10:221–234 Neugebauer V, Galhardo V, Maione S, Mackey SC (2009) Forebrain pain mechanisms. Brain Res Rev 60:226–242 Nosyreva ED, Huber KM (2005) Developmental switch in synaptic mechanisms of hippocampal metabotropic glutamate receptordependent long-term depression. J Neurosci 25:2992–3001 Obara I, Goulding SP, Hu JH, Klugmann M, Worley PF, Szumlinski KK (2013) Nerve injury-induced changes in Homer/glutamate receptor signaling contribute to the development and maintenance of neuropathic pain. Pain 154:1932–1945 Obradovic A, Hwang SM, Scarpa J, Hong SJ, Todorovic SM, JevtovicTodorovic V (2014) CaV3.2 T-Type calcium channels in peripheral sensory neurons are important for mibefradil-induced reversal of hyperalgesia and allodynia in rats with painful diabetic neuropathy. PLoS ONE 9:e91467 Ohkura M, Sasaki T, Sadakari J, Gengyo-Ando K, Kagawa-Nagamura Y, Kobayashi C, Ikegaya Y, Nakai J (2012) Genetically encoded green fluorescent Ca2+ indicators with improved detectability for neuronal Ca2+ signals. PLoS ONE 7:e51286 Oliveira AM, Hemstedt TJ, Bading H (2012) Rescue of aging-associated decline in Dnmt3a2 expression restores cognitive abilities. Nat Neurosci 15:1111–1113 Orestes P, Osuru HP, McIntire WE, Jacus MO, Salajegheh R, Jagodic MM, Choe W, Lee J, Lee SS, Rose KE, Poiro N, Digruccio MR, Krishnan K, Covey DF, Lee JH, Barrett PQ, Jevtovic-Todorovic V, Todorovic SM (2013) Reversal of neuropathic pain in diabetes by targeting glycosylation of Ca (V)3.2 T-type calcium channels. Diabetes 62:3828–3838 Ouyang K, Zheng H, Qin X, Zhang C, Yang D, Wang X, Wu C, Zhou Z, Cheng H (2005) Ca2+ sparks and secretion in dorsal root ganglion neurons. Proc Natl Acad Sci USA 102:12259–12264 Park JS, Voitenko N, Petralia RS, Guan X, Xu JT, Steinberg JP, Takamiya K, Sotnik A, Kopach O, Huganir RL, Tao YX (2009) Persistent inflammation induces GluR2 internalization via NMDA receptortriggered PKC activation in dorsal horn neurons. J Neurosci 29: 3206–3219 Patel R, Bauer CS, Nieto-Rostro M, Margas W, Ferron L, Chaggar K, Crews K, Ramirez JD, Bennett DL, Schwartz A, Dickenson AH, Dolphin AC (2013) alpha2delta-1 gene deletion affects somatosensory neuron function and delays mechanical hypersensitivity in response to peripheral nerve damage. J Neurosci 33: 16412–16426 Pecze L, Blum W, Schwaller B (2013) Mechanism of capsaicin receptor TRPV1-mediated toxicity in pain-sensing neurons focusing on the effects of Na (+)/Ca (2+) fluxes and the Ca (2+)-binding protein calretinin. Biochim Biophys Acta 1833:1680–1691 Penzes P, Cahill ME, Jones KA, Srivastava DP (2008) Convergent CaMK and RacGEF signals control dendritic structure and function. Trends Cell Biol 18:405–413 Persson AK, Black JA, Gasser A, Cheng X, Fischer TZ, Waxman SG (2010) Sodium-calcium exchanger and multiple sodium channel isoforms in intra-epidermal nerve terminals. Mol Pain 6:84 Polgar E, Watanabe M, Hartmann B, Grant SG, Todd AJ (2008) Expression of AMPA receptor subunits at synapses in laminae I-III of the rodent spinal dorsal horn. Mol Pain 4:5

Cell Tissue Res (2014) 357:407–426 Pottorf WJ, Thayer SA (2002) Transient rise in intracellular calcium produces a long-lasting increase in plasma membrane calcium pump activity in rat sensory neurons. J Neurochem 83:1002–1008 Qiu J, Tan YW, Hagenston AM, Martel MA, Kneisel N, Skehel PA, Wyllie DJ, Bading H, Hardingham GE (2013) Mitochondrial calcium uniporter Mcu controls excitotoxicity and is transcriptionally repressed by neuroprotective nuclear calcium signals. Nat Commun 4:2034 Rausell E, Cusick CG, Taub E, Jones EG (1992) Chronic deafferentation in monkeys differentially affects nociceptive and nonnociceptive pathways distinguished by specific calcium-binding proteins and down-regulates gamma-aminobutyric acid type A receptors at thalamic levels. Proc Natl Acad Sci USA 89:2571–2575 Raymond CR, Redman SJ (2006) Spatial segregation of neuronal calcium signals encodes different forms of LTP in rat hippocampus. J Physiol 570:97–111 Rhee SG (2001) Regulation of phosphoinositide-specific phospholipase C. Annu Rev Biochem 70:281–312 Rigaud M, Gemes G, Weyker PD, Cruikshank JM, Kawano T, Wu HE, Hogan QH (2009) Axotomy depletes intracellular calcium stores in primary sensory neurons. Anesthesiology 111:381–392 Rondon LJ, Privat AM, Daulhac L, Davin N, Mazur A, Fialip J, Eschalier A, Courteix C (2010) Magnesium attenuates chronic hypersensitivity and spinal cord NMDA receptor phosphorylation in a rat model of diabetic neuropathic pain. J Physiol 588:4205–4215 Rozov A, Zivkovic AR, Schwarz MK (2012) Homer1 gene products orchestrate Ca (2+)-permeable AMPA receptor distribution and LTP expression. Front Synaptic Neurosci 4:4 Rycroft BK, Vikman KS, Christie MJ (2007) Inflammation reduces the contribution of N-type calcium channels to primary afferent synaptic transmission onto NK1 receptor-positive lamina I neurons in the rat dorsal horn. J Physiol 580:883–894 Sabatini BL, Regehr WG (1997) Control of neurotransmitter release by presynaptic waveform at the granule cell to Purkinje cell synapse. J Neurosci 17:3425–3435 Sala C, Futai K, Yamamoto K, Worley PF, Hayashi Y, Sheng M (2003) Inhibition of dendritic spine morphogenesis and synaptic transmission by activity-inducible protein Homer1a. J Neurosci 23:6327– 6337 Sandkühler J (2009) Models and mechanisms of hyperalgesia and allodynia. Physiol Rev 89:707–758 Saneyoshi T, Fortin DA, Soderling TR (2010) Regulation of spine and synapse formation by activity-dependent intracellular signaling pathways. Curr Opin Neurobiol 20:108–115 Sarantopoulos CD, McCallum JB, Rigaud M, Fuchs A, Kwok WM, Hogan QH (2007) Opposing effects of spinal nerve ligation on calcium-activated potassium currents in axotomized and adjacent mammalian primary afferent neurons. Brain Res 1132:84–99 Schafer DP, Stevens B (2010) Synapse elimination during development and disease: immune molecules take centre stage. Biochem Soc Trans 38:476–481 Schicker KW, Chandaka GK, Geier P, Kubista H, Boehm S (2010) P2Y1 receptors mediate an activation of neuronal calcium-dependent K + channels. J Physiol 588:3713–3725 Schlumm F, Mauceri D, Freitag HE, Bading H (2013) Nuclear calcium signaling regulates nuclear export of a subset of class IIa histone deacetylases following synaptic activity. J Biol Chem 288:8074– 8084 Schwechter B, Tolias KF (2013) Cytoskeletal mechanisms for synaptic potentiation. Commun Integr Biol 6:e27343 Schwechter B, Rosenmund C, Tolias KF (2013) RasGRF2 Rac-GEF activity couples NMDA receptor calcium flux to enhanced synaptic transmission. Proc Natl Acad Sci USA 110:14462–14467 Shi TJ, Xiang Q, Zhang MD, Tortoriello G, Hammarberg H, Mulder J, Fried K, Wagner L, Josephson A, Uhlen M, Harkany T, Hokfelt T (2012) Secretagogin is expressed in sensory CGRP neurons and in

Cell Tissue Res (2014) 357:407–426 spinal cord of mouse and complements other calcium-binding proteins, with a note on rat and human. Mol Pain 8:80 Shishkin V, Potapenko E, Kostyuk E, Girnyk O, Voitenko N, Kostyuk P (2002) Role of mitochondria in intracellular calcium signaling in primary and secondary sensory neurones of rats. Cell Calcium 32: 121–130 Shrivastava AN, Triller A, Sieghart W, Sarto-Jackson I (2011) Regulation of GABA (A) receptor dynamics by interaction with purinergic P2X (2) receptors. J Biol Chem 286:14455–14468 Shutov LP, Kim MS, Houlihan PR, Medvedeva YV, Usachev YM (2013) Mitochondria and plasma membrane Ca2+− ATPase control presynaptic Ca2+ clearance in capsaicin-sensitive rat sensory neurons. J Physiol 591:2443–2462 Simonetti M, Giniatullin R, Fabbretti E (2008) Mechanisms mediating the enhanced gene transcription of P2X3 receptor by calcitonin gene-related peptide in trigeminal sensory neurons. J Biol Chem 283:18743–18752 Simonetti M, Hagenston AM, Vardeh D, Freitag HE, Mauceri D, Lu J, Satagopam VP, Schneider R, Costigan M, Bading H, Kuner R (2013) Nuclear calcium signaling in spinal neurons drives a genomic program required for persistent inflammatory pain. Neuron 77: 43–57 Sokolova E, Nistri A, Giniatullin R (2001) Negative cross talk between anionic GABAA and cationic P2X ionotropic receptors of rat dorsal root ganglion neurons. J Neurosci 21:4958–4968 Song SO, Varner J (2009) Modeling and analysis of the molecular basis of pain in sensory neurons. PLoS ONE 4:e6758 Stone LS, Molliver DC (2009) In search of analgesia: emerging roles of GPCRs in pain. Mol Interv 9:234–251 Study RE, Kral MG (1996) Spontaneous action potential activity in isolated dorsal root ganglion neurons from rats with a painful neuropathy. Pain 65:235–242 Stutzmann GE (2007) The pathogenesis of alzheimers disease is it a lifelong “calciumopathy”? Neuroscientist 13:546–559 Tachibana T, Ogura H, Tokunaga A, Dai Y, Yamanaka H, Seino D, Noguchi K (2004) Plasma membrane calcium ATPase expression in the rat spinal cord. Brain Res Mol Brain Res 131:26–32 Talley EM, Cribbs LL, Lee JH, Daud A, Perez-Reyes E, Bayliss DA (1999) Differential distribution of three members of a gene family encoding low voltage-activated (T-type) calcium channels. J Neurosci 19:1895–1911 Tan AM, Stamboulian S, Chang YW, Zhao P, Hains AB, Waxman SG, Hains BC (2008) Neuropathic pain memory is maintained by Rac1regulated dendritic spine remodeling after spinal cord injury. J Neurosci 28:13173–13183 Tappe A, Klugmann M, Luo C, Hirlinger D, Agarwal N, Benrath J, Ehrengruber MU, During MJ, Kuner R (2006) Synaptic scaffolding protein Homer1a protects against chronic inflammatory pain. Nat Med 12:677–681 Tochiki KK, Cunningham J, Hunt SP, Geranton SM (2012) The expression of spinal methyl-CpG-binding protein 2, DNA methyltransferases and histone deacetylases is modulated in persistent pain states. Mol Pain 8:14 Todorovic SM, Jevtovic-Todorovic V (2011) T-type voltage-gated calcium channels as targets for the development of novel pain therapies. Br J Pharmacol 163:484–495 Todorovic SM, Jevtovic-Todorovic V (2014) Targeting of CaV3.2 T-type calcium channels in peripheral sensory neurons for the treatment of painful diabetic neuropathy. Pflugers Arch 466:701–706 Tong CK, MacDermott AB (2006) Both Ca2 + −permeable and impermeable AMPA receptors contribute to primary synaptic drive onto rat dorsal horn neurons. J Physiol 575:133–144 Torsney C, MacDermott AB (2006) Disinhibition opens the gate to pathological pain signaling in superficial neurokinin 1 receptorexpressing neurons in rat spinal cord. J Neurosci 26:1833–1843

425 Toyoda H, Zhao MG, Ulzhofer B, Wu LJ, Xu H, Seeburg PH, Sprengel R, Kuner R, Zhuo M (2009) Roles of the AMPA receptor subunit GluA1 but not GluA2 in synaptic potentiation and activation of ERK in the anterior cingulate cortex. Mol Pain 5:46 Trang T, Beggs S, Salter MW (2011) Brain-derived neurotrophic factor from microglia: a molecular substrate for neuropathic pain. Neuron Glia Biol 7:99–108 Verkhratsky A, Fernyhough P (2008) Mitochondrial malfunction and Ca2 + dyshomeostasis drive neuronal pathology in diabetes. Cell Calcium 44:112–122 Vikman KS, Rycroft BK, Christie MJ (2008) Switch to Ca2+−permeable AMPA and reduced NR2B NMDA receptor- mediated neurotransmission at dorsal horn nociceptive synapses during inflammatory pain in the rat. J Physiol 586:515–527 von Hehn CA, Baron R, Woolf CJ (2012) Deconstructing the neuropathic pain phenotype to reveal neural mechanisms. Neuron 73:638–652 Wall PD, Devor M (1983) Sensory afferent impulses originate from dorsal root ganglia as well as from the periphery in normal and nerve injured rats. Pain 17:321–339 Wang YY, Wei YY, Huang J, Wang W, Tamamaki N, Li YQ, Wu SX (2009) Expression patterns of 5-HT receptor subtypes 1A and 2A on GABAergic neurons within the spinal dorsal horn of GAD67-GFP knock-in mice. J Chem Neuroanat 38:75–81 Waxman SG, Zamponi GW (2014) Regulating excitability of peripheral afferents: emerging ion channel targets. Nat Neurosci 17:153–163 Weiss N, Black SA, Bladen C, Chen L, Zamponi GW (2013) Surface expression and function of Cav3.2 T-type calcium channels are controlled by asparagine-linked glycosylation. Pflugers Arch 465: 1159–1170 Wen XJ, Li ZJ, Chen ZX, Fang ZY, Yang CX, Li H, Zeng YM (2006) Intrathecal administration of Cav3.2 and Cav3.3 antisense oligonucleotide reverses tactile allodynia and thermal hyperalgesia in rats following chronic compression of dorsal root of ganglion. Acta Pharmacol Sin 27:1547–1552 Wen XJ, Xu SY, Chen ZX, Yang CX, Liang H, Li H (2010) The roles of T-type calcium channel in the development of neuropathic pain following chronic compression of rat dorsal root ganglia. Pharmacology 85:295–300 Werth JL, Usachev YM, Thayer SA (1996) Modulation of calcium efflux from cultured rat dorsal root ganglion neurons. J Neurosci 16:1008– 1015 West AE, Chen WG, Dalva MB, Dolmetsch RE, Kornhauser JM, Shaywitz AJ, Takasu MA, Tao X, Greenberg ME (2001) Calcium regulation of neuronal gene expression. Proc Natl Acad Sci USA 98: 11024–11031 Wiens KM, Lin H, Liao D (2005) Rac1 induces the clustering of AMPA receptors during spinogenesis. J Neurosci 25:10627–10636 Woolf CJ, Ma Q (2007) Nociceptors–noxious stimulus detectors. Neuron 55:353–364 Woolf CJ, Salter MW (2000) Neuronal plasticity: increasing the gain in pain. Science 288:1765–1769 Wu LJ, Zhuo M (2009) Targeting the NMDA receptor subunit NR2B for the treatment of neuropathic pain. Neurotherapeutics 6:693–702 Wu ZZ, Guan BC, Li ZW, Yang Q, Liu CJ, Chen JG (2004) Sustained potentiation by substance P of NMDA-activated current in rat primary sensory neurons. Brain Res 1010:117–126 Wu G, Xie X, Lu ZH, Ledeen RW (2009) Sodium-calcium exchanger complexed with GM1 ganglioside in nuclear membrane transfers calcium from nucleoplasm to endoplasmic reticulum. Proc Natl Acad Sci USA 106:10829–10834 Xiao B, Tu JC, Worley PF (2000) Homer: a link between neural activity and glutamate receptor function. Curr Opin Neurobiol 10: 370–374 Xu Q, Yaksh TL (2011) A brief comparison of the pathophysiology of inflammatory versus neuropathic pain. Curr Opin Anaesthesiol 24: 400–407

426 Xu TL, Li JS, Jin YH, Akaike N (1999) Modulation of the glycine response by Ca2 + −permeable AMPA receptors in rat spinal neurones. J Physiol 514(Pt 3):701–711 Yokoyama K, Kurihara T, Makita K, Tanabe T (2003) Plastic change of N-type Ca channel expression after preconditioning is responsible for prostaglandin E2-induced long-lasting allodynia. Anesthesiology 99:1364–1370 Youn DH, Royle G, Kolaj M, Vissel B, Randic M (2008) Enhanced LTP of primary afferent neurotransmission in AMPA receptor GluR2deficient mice. Pain 136:158–167 Yowtak J, Lee KY, Kim HY, Wang J, Kim HK, Chung K, Chung JM (2011) Reactive oxygen species contribute to neuropathic pain by reducing spinal GABA release. Pain 152:844–852 Yu D, Thakor DK, Han I, Ropper AE, Haragopal H, Sidman RL, Zafonte R, Schachter SC, Teng YD (2013) Alleviation of chronic pain following rat spinal cord compression injury with multimodal actions of huperzine A. Proc Natl Acad Sci USA 110:E746–755 Yue J, Liu L, Liu Z, Shu B, Zhang Y (2013) Upregulation of T-type Ca2+ channels in primary sensory neurons in spinal nerve injury. Spine 38:463–470 Yusaf SP, Goodman J, Gonzalez IM, Bramwell S, Pinnock RD, Dixon AK, Lee K (2001) Streptozocin-induced neuropathy is associated with altered expression of voltage-gated calcium channel subunit mRNAs in rat dorsal root ganglion neurones. Biochem Biophys Res Commun 289:402–406 Zacharova G, Palecek J (2009) Parvalbumin and TRPV1 receptor expression in dorsal root ganglion neurons after acute peripheral inflammation. Physiol Res Acad Sci Bohemoslovaca 58:305–309 Zhang YQ, Gao X, Ji GC, Wu GC (2001) Expression of 5-HT2A receptor mRNA in rat spinal dorsal horn and some nuclei of brainstem after peripheral inflammation. Brain Res 900:146–151

Cell Tissue Res (2014) 357:407–426 Zhang YQ, Gao X, Ji GC, Huang YL, Wu GC, Zhao ZQ (2002) Expression of 5-HT1A receptor mRNA in rat lumbar spinal dorsal horn neurons after peripheral inflammation. Pain 98: 287–295 Zhang Y, Li Y, Yang YR, Zhu HH, Han JS, Wang Y (2007) Distribution of downstream regulatory element antagonist modulator (DREAM) in rat spinal cord and upregulation of its expression during inflammatory pain. Neurochem Res 32:1592–1599 Zhang Z, Cai YQ, Zou F, Bie B, Pan ZZ (2011) Epigenetic suppression of GAD65 expression mediates persistent pain. Nat Med 17:1448– 1455 Zhang XL, Mok LP, Lee KY, Charbonnet M, Gold MS (2012) Inflammation-induced changes in BK (Ca) currents in cutaneous dorsal root ganglion neurons from the adult rat. Mol Pain 8:37 Zhang MD, Tortoriello G, Hsueh B, Tomer R, Ye L, Mitsios N, Borgius L, Grant G, Kiehn O, Watanabe M, Uhlen M, Mulder J, Deisseroth K, Harkany T, Hokfelt TG (2014) Neuronal calcium-binding proteins 1/2 localize to dorsal root ganglia and excitatory spinal neurons and are regulated by nerve injury. Proc Natl Acad Sci USA 111: E1149–1158 Zherebitskaya E, Schapansky J, Akude E, Smith DR, Van der Ploeg R, Solovyova N, Verkhratsky A, Fernyhough P (2012) Sensory neurons derived from diabetic rats have diminished internal Ca2+ stores linked to impaired re-uptake by the endoplasmic reticulum. ASN neuro 4: Zhou L, Huang J, Gao J, Zhang G, Jiang J (2014) NMDA and AMPA receptors in the anterior cingulate cortex mediates visceral pain in visceral hypersensitivity rats. Cell Immunol 287:86–90 Zhuo M, Wu G, Wu LJ (2011) Neuronal and microglial mechanisms of neuropathic pain. Mol Brain 4:31

Neuronal calcium signaling in chronic pain.

Acute physiological pain, the unpleasant sensory response to a noxious stimulus, is essential for animals and humans to avoid potential injury. Pathol...
2MB Sizes 1 Downloads 6 Views