Neuropathic pain represents a significant and mounting burden on patients and society at large. Management of neuropathic pain, however, is both intricate and challenging, exacerbated by the limited quantity and quality of clinically available treatments. On this stage, dysfunctional voltage-gated ion channels, especially the presynaptic N-type voltage-gated calcium channel (VGCC) (Cav2.2) and the tetrodotoxin-sensitive voltage-gated sodium channel (VGSC) (Nav1.7), underlie the pathophysiology of neuropathic pain and serve as high profile therapeutic targets. Indirect regulation of these channels holds promise for the treatment of neuropathic pain. In this review, we focus on collapsin response mediator protein 2 (CRMP2), a protein with emergent roles in voltage-gated ion channel trafficking and discuss the therapeutic potential of targetting this protein.

Pain is universal and ravages effects across the globe without discriminating by age or gender [1]. Reported global sales for analgesics recently surpassed $22 billion, with purchases in the United States alone accounting for $13 billion of that overwhelming total [2]. Existing treatments, however, are inadequate and possess innumerable adverse side effects [3,4]. These therapies are further compromised by limited clinical efficacy and inflexibility across variable pain conditions. For this reason, new paradigms in developing treatments for pain are needed. Here, we focus on collapsin response mediator protein 2 (CRMP2) as a novel therapeutic target uniquely positioned to combat pain via regulation of voltage-gated ion channels.

Amongst various painful neuropathies, hyperexcitability of peripheral sensory neurones and dorsal horn spinal cord neurones represent commonly shared traits [5–7]. The presence of mushroom synapses in these circuits increases, with facilitation of excitatory synaptic transmission and depression of inhibitory input accompanying dendritic spines’ morphological changes to augment pain signal propagation [7,8]. Voltage-gated calcium channels (VGCCs) in these neuronal populations further influence nociceptive signaling by tuning neurotransmitter release and mediating membrane depolarization-dependent intracellular signaling via calcium ions (Ca2+) [9–11]. Notably, opioid receptor-targetted treatments also provide pain relief through downstream modulation of VGCCs [12]. Dysfunction and dysregulation of voltage-gated sodium channels (VGSCs) – namely increased sodium channel activity, current density, and negatively shifted half maximal activation voltage – also contribute to reduced thresholds for action potential initiation and heightened spike firing underlying the pathophysiology of pain [6]. These features highlight the position of primary afferent and second order neurones in neuropathic pain signaling pathways, while simultaneously underscoring the utility of therapeutic approaches targetting VGCCs and VGSCs.

Although several existing pain treatments directly block VGCCs or VGSCs, adverse side effects (such as disruption of normal physiological function in both the central and peripheral nervous systems) limit these drugs’ therapeutic use. Prialt (synthetic ω-conotoxin, also called Ziconotide) is one example of an N-type Ca2+ channel (Cav2.2) specific blocker that has been FDA-approved for clinical treatment of severe pain. However, Ziconotide also inhibits sympathetic norepinephrine release, posing a threat to normal cardiovascular physiology, and so it must be restricted to a narrow therapeutic window and intrathecal route of administration [13,14]. Obstacles have plagued the development of VGSC inhibitors, ranging from non-specificity to blood–brain barrier impermeability, as well. A structurally novel benzodiazepine-based state-dependent Nav1.7 blocker, for example exhibited poor pharmacokinetic properties and inefficacy against inflammatory pain states despite intravenous or intrathecal routes of administration [15–17]. While several other Nav1.7 inhibitors demonstrated high target affinity, low nanomolar affinity for other VGSC isoforms has been an insurmountable hurdle to date [18–20]. Despite these impediments, Ziconotide’s parameters of use and clinical indications validate presynaptic Cav2.2 at the spinal synapse, between primary sensory and spinal cord neurones, as a target for chronic pain. Similarly, the benzodiazepine-based inhibitor’s successful elimination of spontaneous neural firing and reversal of mechanical allodynia, in an experimental rodent model of spinal nerve ligation (SNL)-induced neuropathic pain, reassert the value of Nav1.7 as a therapeutic target [16,17].

Indirectly targetting voltage-gated ion channels through their interacting protein partners represents an emerging approach in drug discovery for neuropathic pain. Recent studies have identified CRMP2 as a part of the nociceptor transcriptome and uncovered CRMP2’s novel interactions with select VGCCs and VGSCs [21–32]. While CRMP2 traditionally promotes neurite outgrowth, neuronal polarization, progenitor proliferation, radial migration and assembly of microtubule networks, interactions along the CRMP2–Cav2.2 and CRMP2–Nav1.7 axes have been increasingly acknowledged and investigated as critical mechanisms mediating pain signals [33–39]. Two variants of CRMP2 exist: a long (~68–75 kDa) and short (~62 kDa) form. Both variants share a common core polypeptide, but divergent N-terminal domains result from alternative mRNA splicing [41]. Here, we focus on the short form, which localizes in both axons and dendrites, while the long form is primarily expressed in neuronal cell bodies [41]. To promote microtubule assembly and neurite outgrowth, CRMP2 enhances tubulin GTPase activity to facilitate tubulin polymerization [40–42]. Interestingly, CRMP2 phosphorylation by cyclin dependent kinase-5 (Cdk5) at Ser522 arrests axonal growth, and expression of both CRMP2 and Cdk5 increase in neuropathic pain states [43–48]. Work demonstrating the capacity of Cdk5-inhibiting compounds, such as roscovitine, to alleviate neuropathic and inflammatory pain provide a link between post-translational modifications of CRMP2 and pain phenotypes [49–51].

In this review, we explore the mechanisms underlying CRMP2’s regulation of VGCC and VGSC surface expression. We also begin to unravel the code of post-translational modifications directing CRMP2 function, especially phosphorylation and SUMOylation (Small Ubiquitin-Like Modifier (SUMO)). By considering contributions of other CRMP2 binding partners, we investigate yet another layer of CRMP2’s exquisite regulation of voltage-gated ion channel trafficking. Finally, we examine the potential of these mechanisms as therapeutic targets for neuropathic pain and implications for on-going drug discovery ventures (Figure 1).

CRMP2 and ion channel signaling

Figure 1
CRMP2 and ion channel signaling

Cav2.2 and Nav1.7 channels exist in at least two pools – membrane-delimited and cytosolic. Cdk5-mediated phosphorylation of CRMP2 (at Ser522) and CRMP2-binding with syntaxin 1A facilitate the CRMP2–Cav2.2 interaction and increased Cav2.2 membrane insertion. In a parallel process, CRMP2 phosphorylation by Cdk5 supports CRMP2 SUMOylation (at Lys374) and subsequent increased Nav1.7 trafficking and surface expression.

Figure 1
CRMP2 and ion channel signaling

Cav2.2 and Nav1.7 channels exist in at least two pools – membrane-delimited and cytosolic. Cdk5-mediated phosphorylation of CRMP2 (at Ser522) and CRMP2-binding with syntaxin 1A facilitate the CRMP2–Cav2.2 interaction and increased Cav2.2 membrane insertion. In a parallel process, CRMP2 phosphorylation by Cdk5 supports CRMP2 SUMOylation (at Lys374) and subsequent increased Nav1.7 trafficking and surface expression.

Close modal

In neurones, Cav2.2 mediates Ca2+ influx as a primary precursor event triggering synaptic vesicle docking [51–55]. Interactions between Cav2.2 and the synaptic protein syntaxin 1A then follow to produce fast, synchronous neurotransmitter release [56–60]. Although multiple naturally occurring splice variants of Cav2.2 exist, a subset of splice forms (i.e. exons e37a and e37b) is highly expressed in presynaptic terminals; notably, neuropathic pain states up-regulate expression of these variants, especially in small diameter nociceptors of dorsal root ganglia (DRG) and within laminae I/II of the dorsal horn of the spinal cord [61,62]. Observations also show that Cav2.2 knockout mice lack typical responses to noxious stimuli, emphasizing these channels’ importance for thermal and mechanical nociception [63–70].

Yet, previous strategies directly blocking the pore-forming α subunit of Cav2.2 yielded limited success, and side effects of Ziconotide, for example ranged from confusion, depression, and hallucinations to decreased alertness, somnolence, orthostatic hypotension, and nausea [71–73]. Gabapentin, which targets the α2δ1 auxiliary subunit of Cav2.2 and dampens α2δ1-dependent VGCC membrane insertion, utilizes an indirect approach and possesses clinical indications for diabetic neuropathy, neuropathic pain, trigeminal neuralgia, and fibromyalgia [74–84]. Recently reported competition between the KCa1.1 (BK) potassium channel’s N-terminus and Cav2.2’s pore-forming α subunit for binding to the α2δ subunit further validates α2δ-targetting for analgesic strategies to combat neuropathic and inflammatory pain [85]. However, long-term use of gabapentin is restricted by evidence for its interference with synaptogenesis, and therefore memory formation, through disruption of α2δ1–thrombospondin interactions [86]. These shortcomings pair with Ziconotide and gabapentin’s state-independent channel blocking properties to mandate extra caution in their clinical applications. Efforts to harness state-dependent mechanisms for modulating Cav2.2 activity have been less risky, cumulating in the development of state-dependent, small molecule inhibitors such as TROX-1 [87,88]. Despite early successes in producing analgesic effects, TROX-1’s off-target interactions with Cav2.1 (P/Q-type channels) and Cav2.3 (R-type channels), combined with induction of mild locomotor and cardiovascular impairment at extremely high doses, have hindered its advancement [87].

Interfering with CRMP2’s calcium channel-binding domain

Indirectly targetting Cav2.2 via its interactions with the adaptor protein CRMP2, however, has proven more fruitful [22–32]. A proteomic screen of Cav2.2 interacting partners identified CRMP2, subsequently, multiple direct CRMP2–Cav2.2 binding interfaces (i.e. several 15-amino acid long CaV-binding domains (CBD1–CBD3)) were discovered using peptide arrays [22,27]. In a series of studies examining targets of synthetic CBD peptides, conjugated to the HIV1 TAT (t-) domain for cell penetrance, only t-CBD3, comprising CRMP2 residues 484–498, succeeded in biochemically inhibiting the CRMP2–Cav2.2 interaction [27]. t-CBD3 also reduced Ca2+ currents in rat DRG sensory neurones and excitatory synaptic transmission in lamina II neurones from spinal cord slices [26]. Presence of t-CBD3 significantly inhibited the frequency of spontaneous excitatory postsynaptic currents (sEPSCs) without altering their amplitude, strongly suggesting that t-CBD3, and therefore CRMP2, functions via presynaptic rather than post-synaptic mechanisms of neurotransmission, likely to be mediated by Cav2.2 [27]. Since CRMP2 facilitates synaptic vesicle loading in hippocampal neurones, inhibition of the CRMP2–Cav2.2 complex would predictably reduce the recruitment of synaptic vesicles to Cav2.2 localized in the membrane [24]. It follows that knockdown of CRMP2 reduced evoked release of the pro-nociceptive neuropeptide calcitonin gene-related peptide (CGRP) in rat DRGs as well [25]. In a parallel strategy, t-CBD3 reduced evoked CGRP release alongside CGRP-dependent dilation of dural blood vessels in rodents; these data provide molecular and cellular underpinnings for t-CBD3’s ability to suppress tactile hypersensitivity in rodent models of HIV treatment-induced peripheral neuropathy and formalin-induced inflammatory pain [27]. Subsequent studies documented t-CBD3-mediated disruption of the CRMP2–Cav2.2 complex and efficacy in migraine, AIDS therapy-induced peripheral neuropathy, and tibial nerve injury-induced neuropathic pain [23,89,90] (Table 1).

Table 1
Biologics, drugs, and genetic strategies targetting CRMP2 in neuropathic pain
Neuropathic pain modelsMolecular effectsReferences
Peptides    
t-CBD3 • Reverses mechanical allodynia in: • Inhibits CRMP2-Cav2.2 interaction [27,89
 - AIDS therapy (ddC)-induced neuropathic pain • Reduces Ca2+ currents and neuronal excitability in DRGs, sEPSCs in lamina II spinal cord slices, evoked CGRP release, and capsaicin- induced meningeal blood flow 
 - 2.5% Formalin-induced pain (but not edema) 
• Reverses capsaicin-induced nocifensive behavior 
• Does not impair locomotion and spatial memory retrieval or induce depression/despair-associated behavior 
t-CBD3-G14F • Reverses mechanical allodynia in: • Reduces capsaicin-stimulated meningeal blood flow [23
 - AIDS therapy (d4T, stavudine)-induced neuropathic pain • Targets trigeminal ganglion neurones 
t-CBD3-A6K • Reverses mechanical allodynia in: • Increases disruption of CRMP2–Cav2.2 interaction (than t-CBD3) [89
 - d4T-induced neuropathic pain • Greater conformational stability (than t-CBD3) 
 • Reduces neuronal excitability, T-type/R-type Ca2+ currents, and evoked CGRP release in DRGs 
 • Reverses mechanical allodynia in d4T-induced neuropathic pain 
R9-CBD3 • Reverses mechanical allodynia in: • Inhibits CRMP2–Cav2.2 interaction [91,93
 - tibial nerve injury-induced neuropathic pain • Reduces depolarization-evoked Ca2+ influx in DRGs 
 - ddC-induced neuropathic pain  
myr-t-CBD3 • Reverses thermal hyperalgesia in: • Higher peptide retention in cellular targets (than t-CBD3) [94
 - carrageenan-induced inflammatory pain • Biochemically inhibits CRMP2–Cav2.2 interaction 
 - paw incision-induced post-surgical pain • Reduces DRG excitability 
• Reverses mechanical allodynia in: • Inhibits Cav2.2 membrane localization, depolarization-evoked Ca2+ influx, and Ca2+ currents 
 - paw incision-induced post-surgical pain 
• Does not impair locomotion, induce paralysis, or cause conditioned place preference for a peptide-paired 
R9-CBD3-A6K • Reversibly attenuates mechanical allodynia in: • Inhibits CRMP2–Cav2.2 interaction and Cav2.2 membrane localization [90
 - tibial nerve injury-induced neuropathic pain • Reduces depolarization-evoked Ca2+ influx in DRGs 
• Does not induce conditioned place preference with a peptide-paired chamber; impair locomotion or learning/recognition memory; induce anxiolytic or depression/despair • Reduces neuronal excitability and T-type/R-type Ca2+ currents in DRGs 
• Induces dopamine release in the nucleus accumbens (in pain states only) 
t-N-terminal fragment-CBD3 • Reverses mechanical allodynia and thermal hyperalgesia in: • Inhibits CRMP2–Cav2.2 interaction, depolarization-evoked Ca2+ influx, and Ca2+ currents in DRGs [92
 - paw incision-induced post-surgical pain  
 - HIV-induced neuropathic pain  
t-CNRP1 • Reverses thermal hyperalgesia in: • Inhibits CRMP2-syntaxin 1A and CRMP2–neurofibromin interactions [115
 - paw incision-induced post-surgical pain • Reduces evoked CGRP release in spinal cord 
 - carrageenan-induced inflammatory pain (but not edema) • Inhibits Cav2.2 membrane localization, depolarization-evoked Ca2+ influx, and Ca2+ currents in DRGs 
• Reverses mechanical allodynia in: 
 - paw incision-induced post-surgical pain 
 - HIV-induced neuropathic pain 
Small molecules    
(S)-Lacosamide • Reverses mechanical allodynia and thermal hyperalgesia in: • Inhibits the CRMP2-Cav2.2 interaction and depolarization-evoked Cav2.2-mediated Ca2+ influx in both DRGs and cortical neurones [101,103,115
 - paw incision-induced post-surgical pain • Directly inhibits Cdk5-mediated CRMP2 phosphorylation in DRGs 
 - SNL-induced neuropathic pain • Inhibits Cav2.2 and Nav1.7 surface expression and Ca2+ currents 
 - SNI-induced neuropathic pain • Reverses CRISPR/Cas9-editing of Nf1 induced hyperexcitability (i.e. increased action potential frequency, reduced rheobase) and changes in Nav1.7/Cav2.2-mediated current density 
 - CRISPR/Cas9-induced neurofibromin truncation (missing C-terminal domain) 
• Does not impair locomotion 
Genetic methods    
CRMP2 overexpression  In hippocampal neurones: [24,25
 • Increases surface Cav2.2 expression and Ca2+ currents 
 • Increases synaptic vesicle recycling 
 • Enhances glutamate release 
 • Increases numbers of synaptic boutons 
 In DRGs: 
 • Increases surface Cav2.2 expression and Ca2+ currents 
CRMP2 shRNA/siRNA/CRISPR • Reverses thermal hyperalgesia in: • Inhibits Ca2+ currents in hippocampal neurones and DRGs [24,25,116
 - t-RFP-Nf1-Cas9 lentivirus-infected rats (missing C-terminal domain of neurofibromin) • Reduces Cav2.2 surface expression, τactivation for Ca2+ currents, and evoked CGRP release in DRGs 
 • Inhibits Na+ currents in DRGs from virus-infected rats 
CBD3 AAV vector • Reverses mechanical allodynia, thermal hyperalgesia, and cold thermal hypersensitivity in: • Inhibits N-type and T-type Ca2+ currents in DRGs from rats injected with AAV-CBD3 [97
 - SNI-induced neuropathic pain  
Neuropathic pain modelsMolecular effectsReferences
Peptides    
t-CBD3 • Reverses mechanical allodynia in: • Inhibits CRMP2-Cav2.2 interaction [27,89
 - AIDS therapy (ddC)-induced neuropathic pain • Reduces Ca2+ currents and neuronal excitability in DRGs, sEPSCs in lamina II spinal cord slices, evoked CGRP release, and capsaicin- induced meningeal blood flow 
 - 2.5% Formalin-induced pain (but not edema) 
• Reverses capsaicin-induced nocifensive behavior 
• Does not impair locomotion and spatial memory retrieval or induce depression/despair-associated behavior 
t-CBD3-G14F • Reverses mechanical allodynia in: • Reduces capsaicin-stimulated meningeal blood flow [23
 - AIDS therapy (d4T, stavudine)-induced neuropathic pain • Targets trigeminal ganglion neurones 
t-CBD3-A6K • Reverses mechanical allodynia in: • Increases disruption of CRMP2–Cav2.2 interaction (than t-CBD3) [89
 - d4T-induced neuropathic pain • Greater conformational stability (than t-CBD3) 
 • Reduces neuronal excitability, T-type/R-type Ca2+ currents, and evoked CGRP release in DRGs 
 • Reverses mechanical allodynia in d4T-induced neuropathic pain 
R9-CBD3 • Reverses mechanical allodynia in: • Inhibits CRMP2–Cav2.2 interaction [91,93
 - tibial nerve injury-induced neuropathic pain • Reduces depolarization-evoked Ca2+ influx in DRGs 
 - ddC-induced neuropathic pain  
myr-t-CBD3 • Reverses thermal hyperalgesia in: • Higher peptide retention in cellular targets (than t-CBD3) [94
 - carrageenan-induced inflammatory pain • Biochemically inhibits CRMP2–Cav2.2 interaction 
 - paw incision-induced post-surgical pain • Reduces DRG excitability 
• Reverses mechanical allodynia in: • Inhibits Cav2.2 membrane localization, depolarization-evoked Ca2+ influx, and Ca2+ currents 
 - paw incision-induced post-surgical pain 
• Does not impair locomotion, induce paralysis, or cause conditioned place preference for a peptide-paired 
R9-CBD3-A6K • Reversibly attenuates mechanical allodynia in: • Inhibits CRMP2–Cav2.2 interaction and Cav2.2 membrane localization [90
 - tibial nerve injury-induced neuropathic pain • Reduces depolarization-evoked Ca2+ influx in DRGs 
• Does not induce conditioned place preference with a peptide-paired chamber; impair locomotion or learning/recognition memory; induce anxiolytic or depression/despair • Reduces neuronal excitability and T-type/R-type Ca2+ currents in DRGs 
• Induces dopamine release in the nucleus accumbens (in pain states only) 
t-N-terminal fragment-CBD3 • Reverses mechanical allodynia and thermal hyperalgesia in: • Inhibits CRMP2–Cav2.2 interaction, depolarization-evoked Ca2+ influx, and Ca2+ currents in DRGs [92
 - paw incision-induced post-surgical pain  
 - HIV-induced neuropathic pain  
t-CNRP1 • Reverses thermal hyperalgesia in: • Inhibits CRMP2-syntaxin 1A and CRMP2–neurofibromin interactions [115
 - paw incision-induced post-surgical pain • Reduces evoked CGRP release in spinal cord 
 - carrageenan-induced inflammatory pain (but not edema) • Inhibits Cav2.2 membrane localization, depolarization-evoked Ca2+ influx, and Ca2+ currents in DRGs 
• Reverses mechanical allodynia in: 
 - paw incision-induced post-surgical pain 
 - HIV-induced neuropathic pain 
Small molecules    
(S)-Lacosamide • Reverses mechanical allodynia and thermal hyperalgesia in: • Inhibits the CRMP2-Cav2.2 interaction and depolarization-evoked Cav2.2-mediated Ca2+ influx in both DRGs and cortical neurones [101,103,115
 - paw incision-induced post-surgical pain • Directly inhibits Cdk5-mediated CRMP2 phosphorylation in DRGs 
 - SNL-induced neuropathic pain • Inhibits Cav2.2 and Nav1.7 surface expression and Ca2+ currents 
 - SNI-induced neuropathic pain • Reverses CRISPR/Cas9-editing of Nf1 induced hyperexcitability (i.e. increased action potential frequency, reduced rheobase) and changes in Nav1.7/Cav2.2-mediated current density 
 - CRISPR/Cas9-induced neurofibromin truncation (missing C-terminal domain) 
• Does not impair locomotion 
Genetic methods    
CRMP2 overexpression  In hippocampal neurones: [24,25
 • Increases surface Cav2.2 expression and Ca2+ currents 
 • Increases synaptic vesicle recycling 
 • Enhances glutamate release 
 • Increases numbers of synaptic boutons 
 In DRGs: 
 • Increases surface Cav2.2 expression and Ca2+ currents 
CRMP2 shRNA/siRNA/CRISPR • Reverses thermal hyperalgesia in: • Inhibits Ca2+ currents in hippocampal neurones and DRGs [24,25,116
 - t-RFP-Nf1-Cas9 lentivirus-infected rats (missing C-terminal domain of neurofibromin) • Reduces Cav2.2 surface expression, τactivation for Ca2+ currents, and evoked CGRP release in DRGs 
 • Inhibits Na+ currents in DRGs from virus-infected rats 
CBD3 AAV vector • Reverses mechanical allodynia, thermal hyperalgesia, and cold thermal hypersensitivity in: • Inhibits N-type and T-type Ca2+ currents in DRGs from rats injected with AAV-CBD3 [97
 - SNI-induced neuropathic pain  

Abbreviations: AAV, adeno-associated viral; CDB3; calcium binding domain 3; CGRP, calcitonin gene-related peptide; CNRP1, CRMP2–neurofibromin regulating peptide; CRISPR, clustered regularly interspaced short palindromic repeats; d4t, 2′,3′-didehydro-2′,3′-dideoxythymidine (Stavudine); ddC, 2’,3’ dideoxycytidine; sEPSC, spontaneous excitatory postsynaptic current; SNI, spared nerve injury.

Importantly, disrupting the CRMP2–Cav2.2 interaction did not alter core sympathetic nervous system functions such as pulsatile arterial pressure, mean arterial pressure, heart rate, and core body temperature [91]. Several evolutionary progressions of the CBD3 peptide were synthesized, with each mutation of the original peptide sequence intended to improve its cell penetrance and structural stability [23,90–94]. Antinociceptive effects of one such peptide were reversible in a rodent model of neuropathic pain, and nociceptive behaviors readily resurfaced once peptide administration ceased [90]. Continuous peptide administration did not result in tolerance to its analgesic properties [90]. Although addiction represents a major hazard of currently available pain therapeutics, CBD3 peptides did not cause conditioned place preference for a peptide administration paired chamber and increased dopamine release in the nucleus accumbens of injured, but not naïve, rodents [90]. These results suggest that uncoupling the CRMP2–Cav2.2 complex is non-addictive, while still engaging the centralized affective component of pain and peripheral sensory mechanisms [95,96]. A battery of assessments for anxiety-, despair/depression-, locomotion-, and cognition/recognition memory-associated effects (i.e. light-dark box, elevated plus maze, open field, tail suspension, rotorod, and novel object tests) also found no significant difference in the behaviors of peptide-treated rodents compared with vehicle-treated counterparts [27,90]. Centralized off-target side effects of CBD3 peptides are unlikely given the wide array of tests demonstrating normal behavior in tasks requiring core central nervous system participation. Providing additional validation for this therapeutic strategy, adenoviral injection of CBD3 into rodents’ lumbar DRGs afforded protection from spared nerve injury (SNI)-induced neuropathic pain for over 6 weeks [97].

Investigating CBD3’s pharmacological mechanism by synthesizing Cav2.2-derived peptides revealed that Cav2.2’s first intracellular loop (L1-Cav2.2, residues 388–402) and distal C-terminus (Ct-Cav2.2, residues 2014–2028) directly bind CRMP2 [97]. Binding affinity of Ct-Cav2.2/CRMP2 was ~75-fold greater than between L1-Cav2.2/CRMP2, and only disruption of the Ct-Cav2.2/CRMP2 interaction inhibited evoked CGRP release in rat DRGs [98]. Notably, there is limited distal Ct homology amongst VGCCs, while sequence conservation of L1 is greater than 60%. In a complementary study to identify CRMP2’s key residues mediating the Cav2.2 interaction, the CBD3 peptide was divided into shorter fragments: the first six amino acids (N-terminal fragment) and the last nine (C-terminal fragment) [93]. In rat DRGs, only CBD3’s N-terminal fragment successfully inhibited depolarization-evoked calcium influx, reduced Ca2+ currents, and disrupted the CRMP2–Cav2.2 interaction in situ [93](Table 1). Through selective single site mutagenesis, further optimization of the N-terminal fragment and in vivo assessments identified three key features facilitating disruption of this interaction by a CBD-derived peptide: (i) a positively charged site corresponding (guanidine group in arginine) at position 4 of the six-mers, (ii) a hydrophobic region (arginine residue) at position 4 of the six-mers; and (iii) a hydrogen bond acceptor (alanine) at the first position within the N-terminal fragment [93]. This discovery establishes three basic pharmacophore elements upon which biochemically and functionally similar small molecules can be designed, harnessing CRMP2’s regulation of Cav2.2 to tackle neuropathic pain.

Cdk5-mediated phosphorylation of CRMP2 regulates Cav2.2 trafficking

The link between phosphorylation of CRMP2 by Cdk5 and pain phenotypes warranted further exploration. Knowledge that Cdk5-mediated phosphorylation of CRMP2 arrests axonal growth encouraged the search for a molecule capable of inhibiting CRMP2-dependent neurite outgrowth. The antiepileptic drug (2R)-2-(acetylamino)-N-benzyl-3-methoxypropanamide ((R)-LCM, tradename Vimpat®) fits this profile, and subsequent characterization of (R)-LCM’s molecular derivatives uncovered (S)-Lacosamide ((S)-LCM), which also targets CRMP2 [43–45,99,100]. (S)-LCM selectively inhibits Cdk5-mediated phosphorylation of CRMP2, and a combination of structural, biophysical studies and in silico docking models contributed to the determination of (S)-LCM’s precise binding pocket [101]. In parallel, it was discovered that phosphorylation of CRMP2 by Cdk5 dynamically regulates and increases the association of CRMP2 with Cav2.2 [102]. While activation of Cdk5 and increased CRMP2 expression in rat DRGs contribute to several pain phenotypes, (S)-LCM biochemically disrupts the CRMP2–Cav2.2 interaction while inhibiting depolarization-evoked Ca2+ influx and Ca2+ currents [46,49,102,103]. Systemic administration of (S)-LCM to mice, at three daily doses of 20 mg/kg over 4 days, successfully uncoupled the CRMP2–Cav2.2 interaction in brain lysates, with no detectable effects on locomotion, feeding, or behavior [103]. In experimental rodent models of SNL- and SNI-induced neuropathic pain, (S)-LCM administration effectively reversed nociceptive behaviors within 30 and 120 min, respectively [103]. These findings add the inhibition of Cdk5-mediated phosphorylation of CRMP2 as a potential therapeutic strategy for eliminating chronic pain.

Unraveling a larger CRMP2 complex: neurofibromin and syntaxin 1a

Studies centered on dysregulation of neurofibromin, a tumor suppressor gene product, have also demonstrated that CRMP2 interacts with neurofibromin’s C-terminal domain, and that the two proteins co-localize in neurites [104,105]. Strikingly, neurofibromatosis type-1 (NF1) patients, in whom neurofibromin is often mutated or truncated, experience idiopathic chronic pain for which opioids often fail to provide relief [106–109]. Observations of DRGs in Nf1+/− haploinsufficient mice revealed augmented sodium (Na+) current densities (both tetrodotoxin (TTX)-sensitive and TTX-resistant), alongside increased N-type Ca2+ currents and enhanced stimulus-evoked release of glutamate, substance P, and CGRP [110–113]. Interestingly, treatment of Nf1+/− DRGs with the aforementioned t-CBD3 inhibited both Cav2.2-mediated currents and CGRP release [98]. These results not only link ion channel dysregulation, but also CRMP2, to the pathophysiology of NF1, and suggest that sensitization of small diameter nociceptive sensory neurones may explain the pain reported by NF1 patients [114].

To investigate the physiological role of the CRMP2–neurofibromin interaction, a CRISPR/Cas9 strategy truncating the C-terminal domain of neurofibromin was adopted and then DRGs from these CRISPR/Cas9-edited rats were electrophysiologically characterized [115]. These recordings demonstrated that CRISPR/Cas9-editing of Nf1 and the resulting neurofibromin truncation in rat DRGs caused a remodeling of ion channels with biophysical properties essentially parallel to those of Nf1+/− DRGs; augmented Na+ and Ca2+ currents were observed in addition to hyperexcitability (determined by increased action potential frequency and decreased threshold), while K+ currents were not significantly different [112]. In both male and female rats, neurofibromin truncation led to the development of thermal hyperalgesia [115], dispeling previous reports of possible sexual dimorphism in the behavioral responses.

Biochemical examination of DRGs showed that neurofibromin truncation permitted an up-regulation in phosphorylation of CRMP2 by Cdk5 [115]. Given the previous identification of (S)-LCM as a selective inhibitor of Cdk5-mediated phosphorylation of CRMP2, DRGs, with neurofibromin truncation were treated with (S)-LCM; this inhibited membrane localization of Cav2.2 and normalized the previously detected changes in both Na+ and Ca2+ currents as well as action potential frequency, threshold, and overall hyperexcitability [115]. Oral administration of (S)-LCM to rats previously subjected to CRISPR/Cas9-editing of Nf1 reduced thermal hyperalgesia in both males and females [115]. A subsequent study demonstrated that lentiviral knockdown of CRMP2 expression in rodents subjected to neurofibromin truncation was sufficient to attenuate thermal hyperalgesia and normalize Na+ current density [116]. Although the precise nature of NF1-linked pain syndromes remains unknown, patients clearly experience pain independent of their tumor burden, and these findings affirm the necessity of CRMP2 in NF1-related pain, while suggesting a physiological role for neurofibromin’s C-terminal domain-mediated inhibition of the CRMP2–Cav2.2 interaction [106].

Efforts to fully understand neurofibromin-mediated regulation of CRMP2, and its consequences for ion channel activity, led to the synthesis of complementary peptide arrays on which CRMP2 and neurofibromin’s C-terminus were tiled, respectively [117]. Multiple binding domains were identified, and corresponding peptides were synthesized – including a unique CRMP2–neurofibromin regulating peptide (CNRP1) that, when conjugated to tat (t-), disrupts the CRMP2–neurofibromin interaction while also inhibiting membrane localization of Cav2.2, depolarization-evoked Ca2+ influx, and Ca2+ currents in rat DRGs [117]. To test the hypothesis that neurofibromin prevents CRMP2’s association with additional protein partners engaged in Cav2.2 trafficking, CRMP2 binding interactions within a nanodisc-embedded synaptic membrane library (which preserved membrane protein integrity in a virtually native environment without use of detergents) were assessed [117]. MS of co-immunoprecipitated CRMP2-bound proteins, in the presence and absence of t-CNRP1, identified syntaxin 1A as a novel CRMP2 binding partner (at residues 456–480) [117].

Syntaxin 1A is widely acknowledged as a regulator of Cav2.2 and Cav2.2-associated neurotransmitter release [54,58,118]. Recruitment of syntaxin 1A to Cav2.2’s intracellular SYNaptic Protein INTeraction (‘synprint’) region facilitates trafficking of the channel and promotes synaptic vesicle docking [55,118–120]. Past efforts to harness uncoupling of the syntaxin 1A–Cav2.2 interaction for therapeutic use have not advanced despite an early report of success [121]. However, these recent discoveries implicating CRMP2 in neurotransmitter release may support reconsideration of this strategy.

Notably, syntaxin 1A binds CRMP2 in the same region as the C-terminal domain of neurofibromin [117]. Disrupting the CRMP2–syntaxin 1A interaction via t-CNRP1 inhibited evoked CGRP release in spinal cord, and intrathecal administration of t-CNRP1 alleviated nociceptive behaviors in carrageenan-induced inflammatory pain, post-surgical pain, and HIV-induced (gp120-evoked) sensory neuropathic pain [117]. This evidence suggests that neurofibromin plays a physiological role in sequestration of CRMP2 from syntaxin 1A to disrupt Cav2.2 trafficking, limit synaptic vesicle docking, mitigate CGRP release, and ultimately, control pain signaling. However, loss of neurofibromin, as observed in NF1 patients, results in disinhibition of the tripartite CRMP2–Cav2.2–syntaxin 1A interaction, therefore permitting increased N-type Ca2+ currents and CGRP release. Precedence for therapeutic targetting of synaptic vesicle proteins already exists (e.g. levetiracetam), and curbing the CRMP2–syntaxin 1A interaction may be especially advantageous in efforts to relieve not only NF1-related pain, but neuropathic pain at large [122].

Synergistic regulation of T-type and R-type VGCCs in neuropathic pain

T-type (i.e. Cav3.1, Cav3.2, and Cav3.3) and R-type (i.e. Cav2.3) Ca2+ channels also represent acknowledged molecular targets for the development of pain therapeutics. In rat DRGs and dorsal horn spinal cord neurones, evidence suggests that both T-type and R-type Ca2+ currents contribute to neuronal excitability [123–126]. While fast kinetic properties and high open probabilities at only very negative membrane potentials characterize T-type Ca2+ channels, R-type Ca2+ channels exhibit slow kinetic properties and, in pyramidal hippocampal neurones, the capacity to modulate both after depolarization and burst firing [127–129]. Previous efforts to harness direct T-type Ca2+ channel blockers (e.g. mibefradil and ethosuximide) yielded mild success in relieving nociceptive behaviors in various neuropathic pain models; antisense oligonucleotide-mediated silencing of T-type channels attenuated nociception in a chronic constrictive injury model of neuropathic pain as well [130–135]. Obstacles posed by side effects and blood–brain barrier impermeability, however, have limited their advancement [130–132]. Likewise, impediments have faced R-type channel blockers – such as SNX-482, which demonstrated antinociceptive effects in several experimental chronic neuropathic pain models, but appears to incompletely block VGSCs in addition to L-type and P/Q-type Ca2+ channels [136–138].

Indirect T-type channel blocking strategies, however, have proven more fruitful and capable of overcoming the aforementioned hurdles. Ubiquitination of Cav3.2 within the III-IV linker, for example controls its surface expression; a homeostatic balance between ubiquitin ligase WWP1 and ubiquitin protease USP5 activity in turn regulate Cav3.2 activity [139]. Given this, one promising strategy disrupts the Cav3.2–USP5 interface with interfering peptides and small molecules that operate on principles similar to that of the CBD3 peptide and its derivatives [140,141]. Uncoupling the Cav3.2–USP5 interaction produces analgesia in vivo, and these advances pave the way for other conceptually similar approaches that focus on a divergent molecular target [140,141].

While it remains unclear whether CRMP2 directly interacts with T-type and R-type Ca2+ channels, previous reports revealed that the CRMP2-derived peptide t-CBD3 reduced T-type Ca2+ currents [89]. A subsequent iteration of the peptide (t-CBD3-A6K, in which the sixth position was mutated from alanine to lysine for improved structural stability) demonstrated similar competence while simultaneously inhibiting T-type and R-type Ca2+ currents [89]. Substitution of the TAT cell penetration motif for a homopolyarginine (R9) strategy (R9-CBD3-A6K) produced a peptide with equivalent functional efficacy and improved membrane permeability [90,93]. R9-CBD3-A6K robustly reversed nociceptive behaviors in an experimental neuropathic pain model without eliciting side effects associated with Ziconotide or TROX-1 [90]. These observations suggest that the CRMP2-derived peptide exploits synergistic inhibition of a broad spectrum of VGCCs to produce potent relief from various neuropathic pain states. The mechanism by which CRMP2 itself mediates these effects remains unknown and requires further investigation. It is clear, however, that CRMP2 lies at a focal point of molecular traffic and represents a central player in the co-ordination of VGCCs. As such, CRMP2 holds substantial value as a therapeutic target for neuropathic pain.

Participation of VGSCs fundamentally shapes the character of action potentials, and variable VGSC-isoform expression profiles impact spike frequency as well as threshold from neurone to neurone [142,143]. Studies integrating clinical evidence with molecular tools have linked numerous mutations in both TTX-sensitive channels (e.g. Nav1.6 and Nav1.7) and TTX-resistant channels (e.g. Nav1.8 and Nav1.9) to inherited neuropathic pain syndromes such as erythromelalgia, paroxysmal extreme pain disorder, and small fiber neuropathy [144–148]. In particular, reports highlighting Nav1.7 enrichment in nociceptors have fostered interest in Nav1.7 as a target in drug discovery for neuropathic pain [149]. Pain states associated with chemotherapy-induced peripheral neuropathy and diabetic neuropathy up-regulate Nav1.7 and Nav1.7-dependent hyperexcitability of primary afferent sensory neurones [6,150,151]. Combined with existing knowledge of VGSC-associated signaling and trafficking molecules, these observations provide a platform for initial investigations, especially through mechanisms interfering with VGSC trafficking and membrane localization. For example, expression of sodium channel β2 subunits, which promote surface expression of TTX-sensitive Na+ channels through interactions with cell adhesion molecules and cytoskeletal proteins, increase in neuropathic pain states [152,153]. However, β2 subunits associate with multiple VGSC isoforms [154,155]. Similarly, the E3 ubiquitin ligase Nedd4-2 (neural precursor cell expressed developmentally down-regulated protein 4) labels VGSCs for endocytosis, but its loss in neuropathic pain states augments both Nav1.7 and Nav1.8 currents [156,157]. While reducing VGSC surface expression through these mechanisms is unlikely to confer specificity, other trafficking-oriented approaches remain and may prove advantageous.

SUMO protein, for instance is known to regulate a variety of voltage-gated ion channels, including Nav1.2, Kv1.5, and Kv2.1, and their membrane insertion [158–160]. SUMOylation is the covalent addition of a ~12-kDa SUMO groups to the lysine residue of a SUMO-interaction motif (SIM, usually a large hydrophobic residue (ψ) before the modification site lysine and a negatively charged amino acid two residues downstream (ψ-K-X-E/D)) [161,162]. The E2 ubiquitin-like ligase and SUMO-conjugating enzyme Ubc9 facilitates addition of SUMO moieties, while SUMO/sentrin-specific peptidases (SENP) 1 and SENP2 enable reversal of SUMOylation by removing them [163,164]. Previous studies have implicated SUMOylation in pain states. The antidiabetic drug rosiglitazone (Avandia®) induces SUMOylation of a peroxisome proliferator-activated receptor to suppress macrophage infiltration and nociceptive behaviors [165]. In rheumatoid arthritis, intrinsically high SUMO-1 levels couple with histone SUMOylation-mediated decreases in SENP1 expression to produce inflammatory pain [166]. Interestingly, amongst TTX-sensitive Na+ channels, only Nav1.7 lacks a putative SUMOylation motif, and its trafficking mechanism had not been clearly delineated [167].

CRMP2 regulates Nav1.7 trafficking

To determine whether SUMOylation of a Nav1.7-binding partner mediates its membrane insertion, sequences of potential Nav1.7 protein partners were scanned for possible SIMs. This spurred the identification of CRMP2’s SIM (KMD, residues 374–376) [28]. In rat DRGs, a SUMO-impaired CRMP2 mutant (i.e. K374A) exhibited decreased Nav1.7-mediated currents and decreased binding in co-immunoprecipitation with Nav1.7, while a biotinylation assay detected significantly reduced surface expression of Nav1.7 [28,29,169]. Notably, Nav1.1- and Nav1.3-mediated currents were unaffected by SUMO-impairment of CRMP2 [28]. Given that an equilibrium naturally exists between CRMP2’s tetrameric and monomeric states, Nav1.7 currents were assessed in CAD cells (derived from a catecholaminergic mouse cell line) transfected with a molar 3:1 ratio of wild-type CRMP2:CRMP2-K374A cDNAs [28,168]. Under these conditions, Nav1.7 currents were markedly reduced in comparison with CAD cells containing only wild-type CRMP2, suggesting that a single SUMO-impaired CRMP2 monomer is sufficient to impact Nav1.7 currents [28]. Although potential CRMP2 SIMs were identified at K20 and K390, mutations at these sites of CRMP2 (i.e. K20A and K390A) did not reduce Nav1.7 currents or Nav1.7 surface expressions [29]. These findings provided the first evidence for CRMP2’s role as a signaling molecule in Nav1.7 trafficking, where loss of CRMP2 SUMOylation drives dissociation of the Nav1.7–CRMP2 interaction.

Subsequent questions centered on regulation of Nav1.7 function by CRMP2’s hierarchical post-translational modification code, including SUMOylation and phosphorylation by Cdk5, Src family tyrosine-kinases Fyn and Yes, Rho-associated protein kinase (RhoK), and glycogen synthase kinase 3β (GSK3β). In rat DRGs, Nav1.7 membrane localization and current density were reduced by a p-null CRMP2 mutant inaccessible to Cdk5 (i.e. S522A), but not by those mutants inaccessible to other kinases (i.e. Fyn (Y32F), Yes (Y479F), GSK-3β (T509A/T514A), RhoK (T555A)) [169]. Notably, impairment of both CRMP2 SUMOylation and Cdk5-mediated phosphorylation (i.e. K374A/S522A) did not further reduce Nav1.7 currents or surface fraction from decreases independently imposed by either mutation [169]. In CAD cells, Cdk5 p-null CRMP2 was inaccessible to SUMOylation, but impairment of CRMP2 SUMOylation did not alter its Cdk5-mediated phosphorylation [163]. Rat DRGs expressing either SUMO-impaired or Cdk5 p-null CRMP2 also exhibited reduced excitability (i.e. decreased action potential frequency) [169]. However, simultaneous loss of CRMP2 SUMOylation and Fyn-mediated phosphorylation (i.e. K374A/Y32F) failed to reduce Nav1.7 currents and surface localization compared with wild-type CRMP2 conditions [169]. Total Nav1.7 expression did not change, and Nav1.1, Nav1.3, Nav1.5, Nav1.6, and TTX-resistant Na+ currents were not affected by impairment of these CRMP2 post-translational modifications [28,169]. Parallel study of Nav1.7 currents in human DRGs revealed congruent effects of SUMO-impaired and Cdk5 p-null CRMP2 mutants on both TTX-sensitive and total Na+ currents [169]. Together, these results convey that Cdk5-mediated phosphorylation is required for CRMP2 SUMOylation and its promotion of Nav1.7 function, while Fyn-mediated phosphorylation is an obligatory precursor to Nav1.7 internalization driven by CRMP2 deSUMOylation.

Previous reports have described Nedd4-2-mediated labeling of Nav1.7 for internalization, in addition to CRMP2-mediated endocytosis of L1-cell adhesion molecules through interactions with the endocytic protein Numb that recruits epidermal growth factor receptor pathway substrate 15 (Eps15) to induce membrane curvature and initiate clathrin-mediated endocytosis [156,157,170–173]. In rat DRGs expressing SUMO-impaired CRMP2, individually eliminating any of these internalization protein partners or pharmacological inhibition of clathrin-mediated endocytosis via Pitstop2, rescued loss of Nav1.7 current [169]. This implies that Numb, Eps15, and Nedd4-2 are requisite participants in Nav1.7’s clathrin-mediated internalization. By contrast, Cdk5-mediated phosphorylation of CRMP2 exhibited reduced interactions with Numb, Eps15, and Nedd4-2 [169]. Not surprisingly, CRMP2 does not bind Numb, Eps15, and Nedd4-2 under loss of Fyn-mediated phosphorylation, regardless of SUMOylation status [169]. This work more completely describes the molecular participants and hierarchy of CRMP2’s post-translational modifications engaged in Nav1.7 internalization. While it remains unclear how CRMP2 exclusively targets Nav1.7 and not other isoforms of VGSCs, examination of intracellular loop sequences unique to Nav1.7 (e.g. the N-terminus and loop 2 connecting the second and third transmembrane domain modules) may reveal additional answers. Regardless, the highlighted conservation of Nav1.7 trafficking mechanisms between rodent and human DRGs presents an ideal opportunity for therapeutic efforts that disrupt the CRMP2–Nav1.7 interaction to selectively impede CRMP2 SUMOylation and attenuate neuropathic pain.

CRMP2: a novel therapeutic target for neuropathic pain

It is apparent that CRMP2 functions as a multidimensional co-ordinator of protein–protein interactions, particularly in mediating surface trafficking of voltage-gated ion channels. Competition between syntaxin 1A and neurofibromin for a singular CRMP2 binding domain regulates CRMP2’s direct interaction with Cav2.2. Cdk5-mediated phosphorylation further enhances the CRMP2–Cav2.2 association. This array of signaling interactions form an attractive target for therapeutic intervention against dysfunctional Cav2.2 in neuropathic pain. Given the potential multipharmacology of CRMP2-oriented strategies that reduce Cav2.2 activity, an obligation exists to explore CRMP2 interactions with T-type Ca2+ channels and R-type Ca2+ channels further. Expansion of the literature on CRMP2-Nav1.7 trafficking mechanisms also reinforces the importance of appreciating how subtle, yet fundamental, molecular events implicate protein in diverse biological roles. CRMP2’s hierarchical post-translational modification code yields immense potential for harnessing block of CRMP2 SUMOylation (as well as other modifications) to reduce Nav1.7 activity and alleviate neuropathic pain without disrupting core physiological functions. Importantly, CRMP2-derived peptides did not demonstrate toxicity or other adverse side effects associated with targetting VGCCs via direct and state-dependent channel blockers. CRMP2-targetted small molecules in development against VGSCs as well as VGCCs ([92,100,101,103,117], unpublished data) have also shown similar profiles of preclinical success. Together, this work precipitates the necessary advancement of CRMP2-centered drug discovery ventures into studies for diverse clinical pain indications and other channelopathies.

The authors declare that there are no competing interests associated with the manuscript.

This work was supported by the Neurofibromatosis New Investigator Award from the Department of Defense Congressionally Directed Military Medical Research and Development Program [grant number NF1000099 to R.K.]; the National Institutes of Health Awards [grant numbers 1R01NS098772, 1R01DA042852 to R.K.]; the Children’s Tumor Foundation NF1 Synodos Award [grant number 2015-04-009A (to R.K.)]; the Funds to the University of Arizona’s Undergraduate Biology Research Program from the University of Arizona’s Senior Vice President for Research’s Office and the Howard Hughes Medical Institute [grant number 52006942 (to L.A.C.)].

USP5

ubiquitin-specific protease 5

BK

big conductance calcium-activated potassium channel

CAD

Catecholaminergic A differentiated

Cdk5

cyclin dependent kinase-5

CGRP

calcitonin gene-related peptide

CNRP1

CRMP2–neurofibromin regulating peptide

CRISPR/Cas9

clustered regularly interspaced short palindromic repeats-associated protein 9- nuclease

CRMP2

collapsin response mediator protein 2

Ct-Cav2.2

distal C-terminus of Cav2.2

DRG

dorsal root ganglia

Eps15

epidermal growth factor receptor pathway substrate 15

FDA

Food and Drug Administration

(S)-LCM

(S)-Lacosamide

L1-Cav2.2

Cav2.2’s first intracellular loop

Nedd4-2

neural precursor cell expressed developmentally down-regulated protein 4

NF1

neurofibromatosis type-1

RhoK

Rho-associated protein kinase

SENP

SUMO/sentrin-specific peptidase

SIM

SUMO-interaction motif

SNI

spared nerve injury

SNL

spinal nerve ligation

SUMO

small ubiquitin-like modifier

TROX-1

N-triazole oxindole

TTX

tetrodotoxin

VGCC

voltage-gated calcium channel

VGSC

voltage-gated sodium channel

WWP1

NEDD4-like E3 ubiquitin-ligase

1
Goldberg
D.S.
and
McGee
S.J.
(
2011
)
Pain as a global public health priority
.
BMC Public Health
11
,
770
[PubMed]
2
Harstall
C.
and
Ospina
M.
(
2003
)
How prevalent is chronic pain?
Pain Clin. Updates
11
,
9
3
Bowersox
S.S.
,
Singh
T.
,
Nadasdi
L.
,
Zukowska-Grojec
Z.
,
Valentino
K.
and
Hoffman
B.B.
(
1992
)
Cardiovascular effects of omega-conopeptides in conscious rats: mechanisms of action
.
J. Cardiovasc. Pharmacol.
20
,
756
764
[PubMed]
4
Wang
Y.X.
,
Bezprozvannaya
S.
,
Bowersox
S.S.
,
Nadasdi
L.
,
Miljanich
G.
,
Mezo
G.
et al
(
1998
)
Peripheral versus central potencies of N-type voltage-sensitive calcium channel blockers
.
Naunyn Schmiedebergs Arch. Pharmacol.
357
,
159
168
[PubMed]
5
Hogan
Q.H.
,
McCallum
J.B.
,
Sarantopoulos
C.
,
Aason
M.
,
Mynlieff
M.
,
Kwok
W.M.
et al
(
2000
)
Painful neuropathy decreases membrane calcium current in mammalian primary afferent neurons
.
Pain
86
,
43
53
[PubMed]
6
Zhang
H.
and
Dougherty
P.M.
(
2014
)
Enhanced excitability of primary sensory neurons and altered gene expression of neuronal ion channels in dorsal root ganglion in paclitaxel- induced peripheral neuropathy
.
Anesthesiology
120
,
1463
1475
[PubMed]
7
Tan
A.M.
,
Stamboulian
S.
,
Chang
Y.W.
,
Zhao
P.
,
Hains
A.B.
,
Waxman
S.G.
et al
(
2008
)
Neuropathic pain memory is maintained by Rac1-regulated dendritic spine remodeling after spinal cord injury
.
J. Neurosci.
28
,
13173
13183
[PubMed]
8
Tan
A.M.
,
Choi
J.S.
,
Waxman
S.G.
and
Hains
B.C.
(
2009
)
Dendritic spine remodeling after spinal cord injury alters neuronal signal processing
.
J. Neurophysiol.
102
,
2396
2409
[PubMed]
9
Bourinet
E.
and
Zamponi
G.W.
(
2005
)
Voltage gated calcium channels as targets for analgesics
.
Curr. Top. Med. Chem.
5
,
539
546
[PubMed]
10
Zamponi
G.W.
,
Feng
Z.P.
,
Zhang
L.
,
Pajouhesh
H.
,
Ding
Y.
and
Belardetti
F.
(
2009
)
Scaffold-based design and synthesis of potent N-type calcium channel blockers
.
Bioorg. Med. Chem. Lett.
19
,
6467
6472
[PubMed]
11
Todorovic
S.M.
and
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
[PubMed]
12
Bourinet
E.
,
Soong
T.W.
,
Stea
A.
and
Snutch
T.P.
(
1996
)
Determinants of the G protein dependent opioid modulation of neuronal calcium channels
.
Proc. Natl. Acad. Sci. U.S.A.
93
,
1486
1491
13
Doggrell
SA.
(
2004
)
Intrathecal ziconotide for refractory pain
.
Expert Opin. Investig. Drugs
13
,
875
877
[PubMed]
14
Wang
Y.X.
,
Bezprozvannaya
S.
,
Bowersox
S.S.
,
Nadasdi
L.
,
Miljanich
G.
,
Mezo
G.
et al
(
2008
)
Peripheral versus central potencies of N-type voltage-sensitive calcium channel blockers
.
Naunyn Schmiedebergs Arch. Pharmacol.
357
,
159
168
15
Theile
J.W.
and
Cummins
T.R.
(
2011
)
Recent developments regarding voltage-gated sodium channel blockers for the treatment of inherited and acquired neuropathic pain syndromes
.
Front. Pharmacol.
2
,
54
[PubMed]
16
Hoyt
S.B.
,
London
C.
,
Gorin
D.
,
Wyvratt
M.J.
,
Fisher
M.H.
,
Abbadie
C.
et al
(
2007
)
Discovery of a novel class of benzazepinone Na(v)1.7 blockers: potential treatments for neuropathic pain
.
Bioorg. Med. Chem. Lett.
17
,
4630
4634
[PubMed]
17
Hoyt
S.B.
,
London
C.
,
Ok
H.
,
Gonzalez
E.
,
Duffy
J.L.
,
Abbadie
C.
et al
(
2007
)
Benzazepinone Nav1.7 blockers: potential treatments for neuropathic pain
.
Bioorg. Med. Chem. Lett.
17
,
6172
6177
[PubMed]
18
Xiao
Y.
,
Blumenthal
K.
,
Jackson
J.O.
II
,
Liang
S.
and
Cummins
T.R.
(
2010
)
The tarantula toxins ProTxII and huwentoxin-IV differentially interact with human Nav1.7 voltage sensors to inhibit channel activation and inactivation
.
Mol. Pharmacol.
78
,
1124
1134
[PubMed]
19
Schmalhofer
W.A.
,
Calhoun
J.
,
Burrows
R.
,
Bailey
T.
,
Kohler
M.G.
,
Weinglass
A.B.
et al
(
2008
)
ProTx-II, a selective inhibitor of NaV1.7 sodium channels, blocks action potential propagation in nociceptors
.
Mol. Pharmacol.
74
,
1476
1484
[PubMed]
20
Lee
J.H.
,
Park
C.K.
,
Chen
G.
,
Han
Q.
,
Xie
R.G.
,
Liu
T.
et al
(
2014
)
A monoclonal antibody that targets a NaV1.7 channel voltage sensor for pain and itch relief
.
Cell
157
,
1393
1404
[PubMed]
21
Thakur
M.
,
Crow
M.
,
Richards
N.
,
Davey
G.I.
,
Levine
E.
,
Kelleher
J.H.
et al
(
2014
)
Defining the nociceptor transcriptome
.
Front. Mol. Neurosci.
7
,
[PubMed]
22
Khanna
R.
,
Zougman
A.
and
Stanley
E.F.
(
2007
)
A proteomic screen for presynaptic terminal N-type calcium channel (CaV2.2) binding partners
.
J. Biochem. Mol. Biol.
40
,
302314
23
Ripsch
M.S.
,
Ballard
C.J.
,
Khanna
M.
,
Hurley
J.H.
,
White
F.A.
and
Khanna
R.
(
2012
)
A peptide uncoupling CRMP-2 from the presynaptic Ca2+ channel complex demonstrate efficacy in animal models of migraine and AIDS therapy-induced neuropathy
.
Transl. Neurosci.
3
,
1
8
[PubMed]
24
Brittain
J.M.
,
Piekarz
A.D.
,
Wang
Y.
,
Kondo
T.
,
Cummins
T.R.
and
Khanna
R.
(
2009
)
An atypical role for collapsin response mediator protein 2 (CRMP-2) in neurotransmitter release via interaction with presynaptic voltage-gated calcium channels
.
J. Biol. Chem.
284
,
31375
31390
[PubMed]
25
Chi
X.X.
,
Schmutzler
B.S.
,
Brittain
J.M.
,
Wang
Y.
,
Hingtgen
C.M.
,
Nicol
G.D.
et al
(
2009
)
Regulation of N-type voltage-gated calcium channels (Cav2.2) and transmitter release by collapsin response mediator protein-2 (CRMP-2) in sensory neurons
.
J. Cell Sci.
122
,
4351
4362
[PubMed]
26
Wilson
S.M.
,
Brittain
J.M.
,
Piekarz
A.D.
,
Ballard
C.J.
,
Ripsch
M.S.
,
Cummins
T.R.
et al
(
2011
)
Further insights into the antinociceptive potential of a peptide disrupting the N-type calcium channel–CRMP-2 signaling complex
.
Channels
5
,
449
456
[PubMed]
27
Brittain
J.M.
,
Duarte
D.B.
,
Wilson
S.M.
,
Zhu
W.
,
Ballard
C.
,
Johnson
P.L.
et al
(
2011
)
Suppression of inflammatory and neuropathic pain by uncoupling CRMP-2 from the presynaptic Ca2+ channel complex
.
Nat. Med.
17
,
822
830
[PubMed]
28
Dustrude
E.T.
,
Wilson
S.M.
,
Ju
W.
,
Xiao
Y.
and
Khanna
R.
(
2013
)
CRMP2 protein SUMOylation modulates NaV1.7 channel trafficking
.
J. Biol. Chem.
288
,
24316
24331
[PubMed]
29
Dustrude
E.T.
,
Perez-Miller
S.
,
Francois-Moutal
L.
,
Moutal
A.
,
Khanna
M.
and
Khanna
R.
(
2017
)
A single structurally conserved SUMOylation site in CRMP2 controls Nav1.7 function
.
Channels
[PubMed]
30
Ju
W.
,
Li
Q.
,
Wilson
S.M.
,
Brittain
J.M.
,
Meroueh
L.
and
Khanna
R.
(
2013
)
SUMOylation alters CRMP2 regulation of calcium influx in sensory neurons
.
Channels
7
,
153
159
[PubMed]
31
Khanna
R.
,
Wilson
S.M.
,
Brittain
J.M.
,
Weimer
J.
,
Sultana
R.
,
Butterfield
A.
et al
(
2012
)
Opening Pandora’s jar: a primer on the putative roles of CRMP2 in a panoply of neurodegenerative, sensory and motor neuron, and central disorders
.
Future Neurol.
7
,
749
771
[PubMed]
32
Feldman
P.
and
Khanna
R.
(
2013
)
Challenging the catechism of therapeutics for chronic neuropathic pain: targeting CaV2.2 interactions with CRMP2 peptides
.
Neurosci. Lett.
557
,
27
36
33
Goshima
Y.
,
Nakamura
F.
,
Strittmatter
P.
and
Strittmatter
S.M.
(
1995
)
Collapsin-induced growth cone collapse mediated by an intracellular protein related to UNC-33
.
Nature
376
,
509
514
[PubMed]
34
Arimura
N.
,
Inagaki
N.
,
Chihara
K.
,
Menager
C.
,
Nakamura
N.
,
Amano
M.
et al
(
2000
)
Phosphorylation of collapsin response mediator protein-2 by Rho-kinase. Evidence for two separate signaling pathways for growth cone collapse
.
J. Biol. Chem.
275
,
23973
23980
[PubMed]
35
Inagaki
N.
,
Chihara
K.
,
Arimura
N.
,
Menager
C.
,
Kawano
Y.
,
Matsuo
N.
et al
(
2001
)
CRMP-2 induces axons in cultured hippocampal neurons
.
Nat. Neurosci.
4
,
781
782
[PubMed]
36
Arimura
N.
,
Menager
C.
,
Fukata
Y.
and
Kaibuchi
K.
(
2004
)
Role of CRMP-2 in neuronal polarity
.
J. Neurobiol.
58
,
34
47
[PubMed]
37
Schmidt
E.F.
and
Strittmatter
S.M.
(
2007
)
The CRMP family of proteins and their role in Sema3A signaling
.
Adv. Exp. Med. Biol.
600
,
1
11
[PubMed]
38
Charrier
E.
,
Reibel
S.
,
Rogemond
V.
,
Aguera
M.
,
Thomasset
N.
and
Honnorat
J.
(
2003
)
Collapsin response mediator proteins (CRMPs): Involvement in nervous system development and adult neurodegenerative disorders
.
Mol. Neurobiol.
28
,
51
64
[PubMed]
39
Hensley
K.
,
Venkova
K.
,
Christov
A.
,
Gunning
W.
and
Park
J.
(
2011
)
Collapsin response mediator protein-2: an emerging pathologic feature and therapeutic target for neurodisease indications
.
Mol. Neurobiol.
43
,
180
191
[PubMed]
40
Yoneda
A.
,
Morgan-Fisher
M.
,
Wait
R.
,
Couchman
J.R.
and
Wewer
U.M.
(
2012
)
A collapsin response mediator protein 2 isoform controls myosin II-mediated cell migration and matrix assembly by trapping ROCK II
.
Mol. Cell. Biol.
32
,
1788
1804
[PubMed]
41
Bretin
S.
,
Reibel
S.
,
Charrier
E.
,
Maus-Moatti
M.
,
Auvergnon
N.
,
Thevenoux
A.
et al
(
2005
)
Differential expression of CRMP1, CRMP2A, CRMP2B, and CRMP5 in axons or dendrites of distinct neurons in the mouse brain
.
J. Comp. Neurol.
486
,
1
17
[PubMed]
42
Chae
Y.C.
,
Lee
S.
,
Heo
K.
,
Ha
S.H.
,
Jung
Y.
,
Kim
J.H.
et al
(
2009
)
Collapsin response mediator protein-2 regulates neurite formation by modulating tubulin GTPase activity
.
Cell. Signal.
29
,
1818
1826
43
Goshima
Y.
,
Hori
H.
,
Sasaki
Y.
,
Yang
T.
,
Kagoshima-Maezono
M.
,
Li
C.
et al
(
1999
)
Growth cone neuropilin-1 mediates collapsin-1/Sema III facilitation of antero- and retrograde axoplasmic transport
.
J. Neurobiol.
39
,
579
589
[PubMed]
44
Uchida
Y.
,
Ohshima
T.
,
Sasaki
Y.
,
Suzuki
H.
,
Yanai
S.
,
Yamashita
N.
et al
(
2005
)
Semaphorin3A signalling is mediated via sequential Cdk5 and GSK3beta phosphorylation of CRMP2: implication of common phosphorylating mechanism underlying axon guidance and Alzheimer’s disease
.
Genes Cells
10
,
165
179
[PubMed]
45
Kamiya
Y.
,
Saeki
K.
,
Takiguchi
M.
and
Funakoshi
K.
(
2013
)
CDK5, CRMP2 and NR2B in spinal dorsal horn and dorsal root ganglion have different role in pain signaling between neuropathic pain model and inflammatory pain model
.
Eur. J. Anaesthesiol.
30
,
214
46
Li
K.
,
Zhao
G.Q.
,
Li
L.Y.
,
Wu
G.Z.
and
Cui
S.S.
(
2014
)
Epigenetic upregulation of Cdk5 in the dorsal horn contributes to neuropathic pain in rats
.
Neuroreport.
25
,
1116
1121
[PubMed]
47
Pareek
T.K.
,
Keller
J.
,
Kesavapany
S.
,
Pant
H.C.
,
Iadarola
M.J.
,
Brady
R.O.
et al
(
2006
)
Cyclin-dependent kinase 5 activity regulates pain signaling
.
Proc. Natl. Acad. Sci. U.S.A.
103
,
791
796
48
Yang
Y.R.
,
He
Y.
,
Zhang
Y.
,
Li
Y.
,
Li
Y.
,
Han
Y.
et al
(
2007
)
Activation of cyclin-dependent kinase 5 (Cdk5) in primary sensory and dorsal horn neurons by peripheral inflammation contributes to heat hyperalgesia
.
Pain
127
,
109
120
[PubMed]
49
Yang
L.
,
Gu
X.
,
Zhang
W.
,
Zhang
J.
and
Ma
Z.
(
2014
)
Cdk5 inhibitor roscovitine alleviates neuropathic pain in the dorsal root ganglia by downregulating N-methyl-D-aspartate receptor subunit 2A
.
Neurol. Sci.
35
,
1365
1371
[PubMed]
50
Yang
Y.R.
,
He
Y.
,
Zhang
Y.
,
Li
Y.
,
Li
Y.
,
Han
Y.
et al
(
2007
)
Activation of cyclin-dependent kinase 5 (Cdk5) in primary sensory and dorsal horn neurons by peripheral inflammation contributes to heat hyperalgesia
.
Pain
127
,
109
120
[PubMed]
51
Catterall
W.A.
and
Few
A.P.
(
2008
)
Calcium channel regulation and presynaptic plasticity
.
Neuron
59
,
882
901
[PubMed]
52
Stanley
E.F.
(
1997
)
The calcium channel and the organization of the presynaptic transmitter release face
.
Trends Neurosci.
20
,
404409
53
Sheng
Z.H.
,
Westenbroek
R.E.
and
Catterall
W.A.
(
1998
)
Physical link and functional coupling of presynaptic calcium channels and the synaptic vesicle docking/fusion machinery
.
J. Bioenerg. Biomembr.
30
,
335
345
[PubMed]
54
Khanna
R.
,
Li
Q.
,
Bewersdorf
J.
and
Stanley
E.F.
(
2007
)
The presynaptic Cav2.2 channel-transmitter release site core complex
.
Eur. J. Neurosci.
26
,
547
559
[PubMed]
55
Mochida
S.
,
Sheng
Z.H.
,
Baker
C.
,
Kobayashi
H.
and
Catterall
W.A.
(
1996
)
Inhibition of neurotransmission by peptides containing the synaptic protein interaction site of N-type Ca2+ channels
.
Neuron
17
,
781
788
[PubMed]
56
Rettig
J.
,
Sheng
Z.H.
,
Kim
D.K.
,
Hodson
C.D.
,
Snutch
T.P.
and
Catterall
W.A.
(
1996
)
Isoform-specific interaction of the alpha1A subunits of brain Ca2+ channels with the presynaptic proteins syntaxin and SNAP-25
.
Proc. Natl Acad. Sci. U.S.A.
93
,
7363
7368
57
Sheng
Z.H.
,
Rettig
J.
,
Cook
T.
and
Catterall
W.A.
(
1996
)
Calcium-dependent interaction of N-type calcium channels with the synaptic core complex
.
Nature
379
,
451
454
[PubMed]
58
Sheng
Z.H.
,
Rettig
J.
,
Takahashi
M.
and
Catterall
W.A.
(
1994
)
Identification of a syntaxin-binding site on N- type calcium channels
.
Neuron
13
,
1303
1313
[PubMed]
59
Bell
T.J.
,
Thaler
C.
,
Castiglioni
A.J.
,
Helton
T.D.
and
Lipscombe
D.
(
2004
)
Cell-specific alternative splicing increases calcium channel current density in the pain pathway
.
Neuron
41
,
127
138
[PubMed]
60
Altier
C.
,
Dale
C.S.
,
Kisilevsky
A.E.
,
Chapman
K.
,
Castiglioni
A.J.
,
Matthews
E.A.
et al
(
2007
)
Differential role of N-type calcium channel splice isoforms in pain
.
J. Neurosci.
27
,
6363
6373
[PubMed]
61
Cizkova
D.
,
Marsala
J.
,
Lukacova
N.
,
Marsala
M.
,
Jergova
S.
,
Orendacova
J.
et al
(
2002
)
Localization of N-type Ca2+ channels in the rat spinal cord following chronic constrictive nerve injury
.
Exp. Brain Res.
147
,
456
463
[PubMed]
62
Westenbroek
R.E.
,
Hell
J.W.
,
Warner
C.
,
Dubel
S.J.
,
Snutch
T.P.
and
Catterall
W.A.
(
1992
)
Biochemical properties and subcellular distribution of an N-type calcium channel alpha 1 subunit
.
Neuron
9
,
1099
1115
[PubMed]
63
Hatakeyama
S.
,
Wakamori
M.
,
Ino
M.
,
Miyamoto
N.
,
Takahashi
E.
,
Yoshinaga
T.
et al
(
2001
)
Differential nociceptive responses in mice lacking the alpha(1B) subunit of N-type Ca(2+) channels
.
Neuroreport
12
,
2423
2427
[PubMed]
64
Kim
C.
,
Jun
K.
,
Lee
T.
,
Kim
S.S.
,
McEnery
M.W.
,
Chin
H.
et al
(
2001
)
Altered nociceptive response in mice deficient in the alpha(1B) subunit of the voltage-dependent calcium channel
.
Mol. Cell. Neurosci.
18
,
235
245
[PubMed]
65
Park
J.
and
Luo
Z.D.
(
2010
)
Calcium channel functions in pain processing
.
Channels (Austin)
4
,
510
517
[PubMed]
66
Zamponi
G.W.
,
Lewis
R.J.
,
Todorovic
S.M.
,
Arneric
S.P.
and
Snutch
T.P.
(
2009
)
Role of voltage-gated calcium channels in ascending pain pathways
.
Brain Res. Rev.
60
,
84
89
[PubMed]
67
Snutch
TP.
(
2005
)
Targeting chronic and neuropathic pain: the N-type calcium channel comes of age
.
NeuroRx.
2
,
662
670
[PubMed]
68
Malmberg
A.B.
and
Yaksh
T.L.
(
1994
)
Voltage-sensitive calcium channels in spinal nociceptive processing. Blockade of N- and P-type channels inhibits formalin-induced nociception
.
J. Neurosci.
14
,
4882
4890
[PubMed]
69
Saegusa
H.
,
Kurihara
T.
,
Zong
S.
,
Kazuno
A.
,
Matsuda
Y.
,
Nonaka
T.
et al
(
2001
)
Suppression of inflammatory and neuropathic pain symptoms in mice lacking the N-type Ca2+ channel
.
EMBO J.
20
,
2349
2356
[PubMed]
70
Callaghan
B.
and
Adams
D.J.
(
2010
)
Analgesic alpha-conotoxins Vc1.1 and RgIA inhibit N-type calcium channels in sensory neurons of alpha9 nicotinic receptor knockout mice
.
Channels (Austin)
4
,
51
54
[PubMed]
71
Thompson
J.C.
,
Dunbar
E.
and
Laye
R.R.
(
2006
)
Treatment challenges and complications with ziconotide monotherapy in established pump patients
.
Pain Physician
9
,
147
152
[PubMed]
72
Skov
M.J.
,
Beck
J.C.
,
de Kater
A.W.
and
Shopp
G.M.
(
2007
)
Nonclinical safety of ziconotide: an intrathecal analgesic of a new pharmaceutical class
.
Int. J. Toxicol.
26
,
411
421
[PubMed]
73
Rauck
R.L.
,
Wallace
M.S.
,
Burton
A.W.
,
Kapural
L.
and
North
J.M.
(
2009
)
Intrathecal ziconotide for neuropathic pain: a review
.
Pain Pract.
9
,
327
337
[PubMed]
74
Morello
C.M.
,
Leckband
S.G.
,
Stoner
C.P.
,
Moorhouse
D.F.
and
Sahagian
G.A.
(
1999
)
Randomized double-blind study comparing the efficacy of gabapentin with amitriptyline on diabetic peripheral neuropathy pain
.
Arch. Intern. Med.
159
,
1931
1937
[PubMed]
75
Gorson
K.C.
,
Schott
C.
,
Herman
R.
,
Ropper
A.H.
and
Rand
W.M.
(
1999
)
Gabapentin in the treatment of painful diabetic neuropathy: a placebo controlled, double blind, crossover trial
.
J. Neurol. Neurosurg. Psychiatry
66
,
251
252
[PubMed]
76
Chong
M.S.
and
Hester
J.
(
2007
)
Diabetic painful neuropathy: current and future treatment options
.
Drugs
67
,
569
585
[PubMed]
77
Backonja
M.
,
Beydoun
A.
,
Edwards
K.R.
,
Schwartz
S.L.
,
Fonseca
V.
,
Hes
M.
et al
(
1998
)
Gabapentin for the symptomatic treatment of painful neuropathy in patients with diabetes mellitus: a randomized controlled trial
.
JAMA
280
,
1831
1836
78
Ahn
S.H.
,
Park
H.W.
,
Lee
B.S.
,
Moon
H.W.
,
Jang
S.H.
,
Sakong
J.
et al
(
2003
)
Gabapentin effect on neuropathic pain compared among patients with spinal cord injury and different durations of symptoms
.
Spine
28
,
341
346
,
[PubMed]
79
Bosnjak
S.
,
Jelic
S.
,
Susnjar
S.
and
Luki
V.
(
2002
)
Gabapentin for relief of neuropathic pain related to anticancer treatment: a preliminary study
.
J. Chemother.
14
,
214
219
[PubMed]
80
Serpell
MG.
(
2002
)
Gabapentin in neuropathic pain syndromes: a randomised, double-blind, placebo- controlled trial
.
Pain
99
,
557
566
[PubMed]
81
Pexton
T.
,
Moeller-Bertram
T.
,
Schilling
J.M.
and
Wallace
M.S.
(
2011
)
Targeting voltage-gated calcium channels for the treatment of neuropathic pain: a review of drug development
.
Expert Opin. Investig. Drugs
20
,
1277
1284
[PubMed]
82
Moore
R.A.
,
Wiffen
P.J.
,
Derry
S.
and
McQuay
H.J.
(
2011
)
Gabapentin for chronic neuropathic pain and fibromyalgia in adults
.
Cochrane Database Syst. Rev.
,
CD007938
,
83
Dolphin
A.C.
(
2012
)
Calcium channel auxiliary α2δ and β subunits: trafficking and one step beyond
.
Nat. Rev. Neurosci.
13
,
542
[PubMed]
84
Zareba
G.
(
2008
)
New treatment options in the management of fibromyalgia: role of pregabalin
.
Neuropsychiatr. Dis. Treat.
4
,
1193
1201
[PubMed]
85
Zhang
F.X.
,
Gadotti
V.M.
,
Souza
I.A.
,
Chen
L.
and
Zamponi
G.W.
(
2018
)
BK potassium channels suppress Cavα2δ subunit function to reduce inflammatory and neuropathic pain
.
Cell Rep.
22
,
1956
1964
[PubMed]
86
Eroglu
C.
,
Allen
N.J.
,
Susman
M.W.
,
O’Rourke
N.A.
,
Park
C.Y.
,
Ozkan
E.
et al
(
2009
)
Gabapentin receptor alpha2delta-1 is a neuronal thrombospondin receptor responsible for excitatory CNS synaptogenesis
.
Cell
139
,
380
392
[PubMed]
87
Abbadie
C.
,
McManus
O.B.
,
Sun
S.Y.
,
Bugianesi
R.M.
,
Dai
G.
,
Haedo
R.J.
et al
(
2010
)
Analgesic effects of a substituted N-triazole oxindole (TROX-1), a state-dependent, voltage-gated calcium channel 2 blocker
.
J. Pharmacol. Exp. Ther.
334
,
545
555
[PubMed]
88
Swensen
A.M.
,
Herrington
J.
,
Bugianesi
R.M.
,
Dai
G.
,
Haedo
R.J.
,
Ratliff
K.S.
et al
(
2011
)
Characterization of the substituted N-triazole oxindole, TROX-1, a small molecule, state-dependent inhibitor of CaV2 calcium channels
.
Mol. Pharmacol.
81
,
488
497
[PubMed]
89
Piekarz
A.D.
,
Due
M.R.
,
Khanna
M.
,
Wang
B.
,
Ripsch
M.S.
,
Wang
R.
et al
(
2012
)
CRMP-2 peptide mediated decrease of high and low voltage-activated calcium channels, attenuation of nociceptor excitability, and antinociception in a model of AIDS therapy-induced painful peripheral neuropathy
.
Mol. Pain
8
,
54
[PubMed]
90
Xie
J.Y.
,
Chew
L.A.
,
Yang
X.
,
Wang
Y.
,
Qu
C.
,
Wang
W.Y.
et al
(
2016
)
Sustained relief of ongoing experimental neuropathic pain by a CRMP2 peptide aptamer with low abuse potential
.
Pain
157
,
2124
2140
[PubMed]
91
Ju
W.
,
Li
Q.
,
Allette
Y.M.
,
Ripsch
M.S.
,
White
F.A.
and
Khanna
R.
(
2013
)
Suppression of pain-related behavior in two distinct rodent models of peripheral neuropathy by a homopolyarginine-conjugated CRMP2 peptide
.
J. Neurochem.
124
,
869
879
[PubMed]
92
Moutal
A.
,
Li
W.
,
Wang
Y.
,
Ju
W.
,
Luo
S.
,
Cai
S.
et al
(
2017
)
Homology-guided mutational analysis reveals the functional requirements for antinociceptive specificity of collapsin response mediator protein 2-derived peptides
.
Br. J. Pharmacol.
,
[PubMed]
93
Moutal
A.
,
Francois-Moutal
L.
,
Brittain
J.M.
,
Khanna
M.
and
Khanna
R.
(
2015
)
Differential neuroprotective potential of CRMP2 peptide aptamers conjugated to cationic, hydrophobic, and amphipathic cell penetrating peptides
.
Front. Cell Neurosci.
8
,
[PubMed]
94
François-Moutal
L.
,
Wang
Y.
,
Moutal
A.
,
Cottier
K.E.
,
Melemedjian
O.K.
,
Yang
X.
et al
(
2015
)
A membrane-delimited N-myristoylated CRMP2 peptide aptamer inhibits CaV2.2 trafficking and reverses inflammatory and post-operative pain behaviors
.
Pain
156
,
1247
1264
[PubMed]
95
Navratilova
E.
,
Xie
J.Y.
,
Okun
A.
,
Qu
C.
,
Eyde
N.
,
Ci
S.
et al
(
2012
)
Pain relief produces negative reinforcement through activation of mesolimbic reward-valuation circuitry
.
Proc. Natl Acad. Sci. U.S.A.
109
,
20709
20713
96
Xie
J.Y.
,
Qu
C.
,
Patwardhan
A.
,
Ossipov
M.H.
,
Navratilova
E.
,
Becerra
L.
et al
(
2014
)
Activation of mesocorticolimbic reward circuits for assessment of relief of ongoing pain: a potential biomarker of efficacy
.
Pain
155
,
1659
1666
[PubMed]
97
Fischer
G.
,
Pan
B.
,
Vilceanu
D.
,
Hogan
Q.H.
and
Yu
H.
(
2014
)
Sustained relief of neuropathic pain by AAV-targeted expression of CBD3 peptide in rat dorsal root ganglion
.
Gene Ther.
21
,
44
51
[PubMed]
98
Wilson
S.M.
,
Schmutzler
B.S.
,
Brittain
J.M.
,
Dustrude
E.T.
,
Ripsch
M.S.
,
Pellman
J.J.
et al
(
2012
)
Inhibition of transmitter release and attenuation of anti-retroviral-associated and tibial nerve injury-related painful peripheral neuropathy by novel synthetic Ca2+ channel peptides
.
J. Biol. Chem.
287
,
35065
35077
[PubMed]
99
Wilson
S.M.
,
Xiong
W.
,
Wang
Y.
,
Ping
X.
,
Head
J.D.
,
Brittain
J.M.
et al
(
2012
)
Prevention of post-traumatic axon sprouting by blocking collapsin response mediator protein 2-mediated neurite outgrowth and tubulin polymerization
.
Neurosci.
210
,
451
466
100
Wilson
S.M.
,
Moutal
A.
,
Melemedjian
O.K.
,
Wang
Y.
,
Ju
W.
,
Francois-Moutal
L.
et al
(
2014
)
The functionalized amino acid (S)-Lacosamide subverts CRMP2-mediated tubulin polymerization to prevent constitutive and activity-dependent increase in neurite outgrowth
.
Front. Cell Neurosci.
8
,
196
[PubMed]
101
Moutal
A.
,
François-Moutal
L.
,
Perez-Miller
S.
,
Cottier
K.
,
Chew
L.A.
,
Yeon
S.K.
et al
(
2015
)
(S)-Lacosamide binding to collapsin response mediator protein 2 (CRMP2) regulates CaV2. 2 activity by subverting its phosphorylation by Cdk5
.
Mol. Neurobiol.
53
,
1959
1976
102
Brittain
J.M.
,
Wang
Y.
,
Eruvwetere
O.
and
Khanna
R.
(
2012
)
Cdk5-mediated phosphorylation of CRMP-2 enhances its interaction with CaV2.2
.
FEBS Lett.
586
,
3813
3818
[PubMed]
103
Moutal
A.
,
Chew
L.A.
,
Yang
X.
,
Wang
Y.
,
Yeon
S.K.
,
Telemi
E.
et al
(
2016
)
(S)-Lacosamide inhibition of CRMP2 phosphorylation reduces postoperative and neuropathic pain behaviors through distinct classes of sensory neurons identified by constellation pharmacology
.
Pain
104
Patrakitkomjorn
S.
,
Kobayashi
D.
,
Morikawa
T.
,
Wilson
M.M.
,
Tsubota
N.
,
Irie
A.
et al
(
2008
)
Neurofibromatosis type 1 (NF1) tumor suppressor, neurofibromin, regulates the neuronal differentiation of PC12 cells via its associating protein, CRMP-2
.
J. Biol. Chem.
283
,
9399
9413
[PubMed]
105
Lin
Y.L.
and
Hsueh
Y.P.
(
2008
)
Neurofibromin interacts with CRMP-2 and CRMP-4 in rat brain
.
Biochem. Biophys. Res. Commun.
369
,
747
752
[PubMed]
106
Creange
A.
,
Zeller
J.
,
Rostaing-Rigattieri
S.
,
Brugieres
P.
,
Degos
J.D.
,
Revuz
J.
et al
(
1999
)
Neurological complications of neurofibromatosis type 1 in adulthood
.
Brain
122
,
473
481
[PubMed]
107
Drouet
A.
,
Wolkenstein
P.
,
Lefaucheur
J.P.
,
Pinson
S.
,
Combemale
P.
,
Gherardi
R.K.
et al
(
2004
)
Neurofibromatosis 1-associated neuropathies: a reappraisal
.
Brain
127
,
1993
2009
[PubMed]
108
Peron
S.M.
and
Baldwin
A.
(
2016
)
Pain in Neurofibromatosis-1
,
Neurofibromatosis Midwest
,
St. Charles, IL
109
Esposito
T.
,
Piluso
G.
,
Saracino
D.
,
Uccello
R.
,
Schettino
C.
,
Dato
C.
et al
(
2015
)
A novel diagnostic method to detect truncated neurofibromin in neurofibromatosis 1
.
J. Neurochem.
135
,
1123
1128
[PubMed]
110
Wang
Y.
,
Duan
J.H.
,
Hingtgen
C.M.
and
Nicol
G.D.
(
2010
)
Augmented sodium currents contribute to enhanced excitability of small diameter capsaicin-sensitive Nf1+/ − mouse sensory neurons
.
J. Neurophysiol.
103
,
2085
2094
[PubMed]
111
Hingtgen
C.M.
,
Roy
S.L.
and
Clapp
D.W.
(
2006
)
Stimulus-evoked release of neuropeptides is enhanced in sensory neurons from mice with a heterozygous mutation of the Nf1 gene
.
Neurosci
137
,
637
645
112
Wang
Y.
,
Brittain
J.M.
,
Wilson
S.M.
,
Hingtgen
C.M.
and
Khanna
R.
(
2010
)
Altered calcium currents and axonal growth in Nf1 haploinsufficient mice
.
Transl. Neurosci.
1
,
106
114
[PubMed]
113
Duan
J.H.
,
Hodgdon
K.E.
,
Hingtgen
C.M.
and
Nicol
G.D.
(
2014
)
N-type calcium current, Cav2.2, is enhanced in small-diameter sensory neurons isolated from Nf1+/ − mice
.
Neurosci.
270
,
192
202
114
Moutal
A.
,
Dustrude
E.T.
and
Khanna
R.
(
2017
)
Sensitization of ion channels contributes to central and peripheral dysfunction in neurofibromatosis type 1
.
Mol. Neurobiol.
54
,
3342
3349
[PubMed]
115
Moutal
A.
,
Yang
X.
,
Li
W.
,
Gilbraith
K.B.
,
Luo
S.
,
Cai
S.
et al
(
2017
)
CRISPR/Cas9 editing of Nf1 gene identifies CRMP2 as a therapeutic target in neurofibromatosis type 1 (NF1)-related pain that is reversed by (S)-Lacosamide
.
Pain
,
116
Moutal
A.
,
Cai
S.
,
Luo
S.
,
Voisin
R.
and
Khanna
R.
(
2017
)
CRMP2 is necessary for neurofibromatosis type 1 related pain
.
Channels
,
12
,
47
50
,
[PubMed]
117
Moutal
A.
,
Wang
Y.
,
Yang
X.
,
Ji
Y.
,
Luo
S.
,
Dorame
A.
et al
(
2017
)
Dissecting the role of the CRMP2-Neurofibromin complex on pain behaviors
.
Pain
158
,
2203
2221
[PubMed]
118
Harkins
A.B.
,
Cahill
A.L.
,
Powers
J.F.
,
Tischler
A.S.
and
Fox
A.P.
(
2004
)
Deletion of the synaptic protein interaction site of the N-type (CaV2.2) calcium channel inhibits secretion in mouse pheochromocytoma cells
.
Proc. Natl Acad. Sci. U.S.A.
101
,
15219
15224
119
Bennett
M.K.
,
Calakos
N.
and
Scheller
R.H.
(
1992
)
Syntaxin: a synaptic protein implicated in docking of synaptic vesicles at presynaptic active zones
.
Science
257
,
255
259
[PubMed]
120
Seagar
M.
and
Takahashi
M.
(
1998
)
Interactions between presynaptic calcium channels and proteins implicated in synaptic vesicle trafficking and exocytosis
.
J. Bioenerg. Biomembr.
30
,
347
356
[PubMed]
121
Catterall
W.
and
Sheng
Z.H.
(
1996
)
Methods and compositions for screening for presynaptic calcium channel blockers
,
U.S. Pat. US08337602
122
Lynch
B.A.
,
Lambeng
N.
,
Nocka
K.
,
Kensel-Hammes
P.
,
Bajjalieh
S.M.
,
Matagne
A.
et al
(
2004
)
The synaptic vesicle protein SV2A is the binding site for the antiepileptic drug levetiracetam
.
Proc. Natl Acad. Sci. U.S.A.
101
,
9861
9866
123
Scroggs
R.S.
and
Fox
A.P.
(
1992
)
Calcium current variation between acutely isolated adult rat dorsal root ganglion neurons of different size
.
J. Physiol.
445
,
639
658
[PubMed]
124
Nelson
M.T.
,
Joksovic
P.M.
,
Perez-Reyes
E.
and
Todorovic
S.M.
(
2005
)
The endogenous redox agent L-cysteine induces T-type Ca2+ channel-dependent sensitization of a novel subpopulation of rat peripheral nociceptors
.
J. Neurosci.
25
,
8766
8775
[PubMed]
125
Cain
S.M.
and
Snutch
T.P.
(
2010
)
Contributions of T-type calcium channel isoforms to neuronal firing
.
Channels (Austin)
4
,
475
482
[PubMed]
126
Fang
Z.
,
Park
C.K.
,
Li
H.Y.
,
Kim
H.Y.
,
Park
S.H.
,
Jung
S.J.
et al
(
2007
)
Molecular basis of Ca(v)2.3 calcium channels in rat nociceptive neurons
.
J. Biol. Chem.
282
,
4757
4764
[PubMed]
127
Chevalier
M.
,
Lory
P.
,
Mironneau
C.
,
Macrez
N.
and
Quignard
J.F.
(
2006
)
T-type CaV3.3 calcium channels produce spontaneous low-threshold action potentials and intracellular calcium oscillations
.
Eur. J. Neurosci.
23
,
2321
2329
[PubMed]
128
Nelson
M.T.
,
Todorovic
S.M.
and
Perez-Reyes
E.
(
2006
)
The role of T-type calcium channels in epilepsy and pain
.
Curr. Pharm. Des.
12
,
2189
2197
[PubMed]
129
Metz
A.E.
,
Jarsky
T.
,
Martina
M.
and
Spruston
N.
(
2005
)
R-type calcium channels contribute to afterdepolarization and bursting in hippocampal CA1 pyramidal neurons
.
J. Neurosci.
25
,
5763
5773
[PubMed]
130
Clozel
J.P.
,
Ertel
E.A.
and
Ertel
S.I.
(
1997
)
Discovery and main pharmacological properties of mibefradil (Ro 40–5967), the first selective T-type calcium channel blocker
.
J. Hypertens. Suppl.
15
,
S17
S25
[PubMed]
131
Ertel
S.I.
and
Clozel
J.P.
(
1997
)
Mibefradil (Ro 40–5967): the first selective T-type Ca2+ channel blocker
.
Expert Opin. Investig. Drugs
6
,
569
582
[PubMed]
132
Dogrul
A.
,
Gardell
L.R.
,
Ossipov
M.H.
,
Tulunay
F.C.
,
Lai
J.
and
Porreca
F.
(
2003
)
Reversal of experimental neuropathic pain by T-type calcium channel blockers
.
Pain
105
,
159
168
[PubMed]
133
Messinger
R.B.
,
Naik
A.K.
,
Jagodic
M.M.
,
Nelson
M.T.
,
Lee
W.Y.
,
Choe
W.J.
et al
(
2009
)
In vivo silencing of the Ca(V)3.2 T-type calcium channels in sensory neurons alleviates hyperalgesia in rats with streptozocin-induced diabetic neuropathy
.
Pain
145
,
184
195
[PubMed]
134
Jagodic
M.M.
,
Pathirathna
S.
,
Joksovic
P.M.
,
Lee
W.
,
Nelson
M.T.
,
Naik
A.K.
et al
(
2008
)
Upregulation of the T-type calcium current in small rat sensory neurons after chronic constrictive injury of the sciatic nerve
.
J. Neurophysiol.
99
,
3151
3156
[PubMed]
135
Wen
X.J.
,
Xu
S.Y.
,
Chen
Z.X.
,
Yang
C.X.
,
Liang
H.
and
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
[PubMed]
136
Newcomb
R.
,
Szoke
B.
,
Palma
A.
,
Wang
G.
,
Chen
X.H.
,
Hopkins
W.
et al
(
1998
)
Selective peptide antagonist of the class E calcium channel from the venom of the tarantula Hysterocrates gigas
.
Biochemistry
37
,
15353
15362
[PubMed]
137
Bourinet
E.
,
Stotz
S.C.
,
Spaetgens
R.L.
,
Dayanithi
G.
,
Lemos
J.
,
Nargeot
J.
et al
(
2001
)
Interaction of SNX482 with domains III and IV inhibits activation gating of alpha(1E) (Ca(V)2.3) calcium channels
.
Biophys. J.
81
,
79
88
[PubMed]
138
Arroyo
G.
,
Aldea
M.
,
Fuentealba
J.
,
Albillos
A.
and
García
A.G.
(
2003
)
SNX482 selectively blocks P/Q Ca2+ channels and delays the inactivation of Na+ channels of chromaffin cells
.
Eur. J. Pharmacol.
475
,
11
18
[PubMed]
139
García-Caballero
A.
,
Gadotti
V.M.
,
Stemkowski
P.
,
Weiss
N.
,
Souza
I.A.
,
Hodgkinson
V.
et al
(
2014
)
The deubiquitinating enzyme USP5 modulates neuropathic and inflammatory pain by enhancing Cav3.2 channel activity
.
Neuron
83
,
1144
1158
[PubMed]
140
Garcia-Caballero
A.
,
Gadotti
V.M.
,
Chen
L.
and
Zamponi
G.W.
(
2016
)
A cell-permeant peptide corresponding to the cUBP domain of USP5 reverses inflammatory and neuropathic pain
.
Mol. Pain
141
Gadotti
V.M.
,
Caballero
A.G.
,
Berger
N.D.
,
Gladding
C.M.
,
Chen
L.
,
Pfeifer
T.A.
et al
(
2015
)
Small organic molecule disruptors of Cav3.2 – USP5 interactions reverse inflammatory and neuropathic pain
.
Mol. Pain
11
,
12
[PubMed]
142
Cummins
T.R.
and
Rush
A.M.
(
2007
)
Voltage-gated sodium channel blockers for the treatment of neuropathic pain
.
Expert Rev. Neurother.
7
,
1597
1612
[PubMed]
143
Yin
R.
,
Liu
D.
,
Chhoa
M.
,
Li
C.M.
,
Luo
Y.
,
Zhang
M.
et al
(
2016
)
Voltage-gated sodium channel function and expression in injured and uninjured rat dorsal root ganglia neurons
.
Int. J. Neurosci.
126
,
182
192
[PubMed]
144
Jarecki
B.W.
,
Sheets
P.L.
,
Jackson
J.O.
II
and
Cummins
T.R.
(
2008
)
Paroxysmal extreme pain disorder mutations within the D3/S4-S5 linker of Nav1.7 cause moderate destabilization of fast inactivation
.
J. Physiol.
586
,
4137
4153
[PubMed]
145
Yang
Y.
,
Huang
J.
,
Mis
M.A.
,
Estacion
M.
,
Macala
L.
,
Shah
P.
et al
(
2016
)
Nav1.7-A1632G mutation from a family with inherited erythromelalgia: enhanced firing of dorsal root ganglia neurons evoked by thermal stimuli
.
J. Neurosci.
36
,
7511
7522
[PubMed]
146
Tanaka
B.S.
,
Zhao
P.
,
Dib-Hajj
F.B.
,
Morisset
V.
,
Tate
S.
,
Waxman
S.G.
et al
(
2016
)
A gain-of-function mutation in Nav1.6 in a case of trigeminal neuralgia
.
Mol. Med.
[PubMed]
147
Han
C.
,
Yang
Y.
,
Te Morsche
R.H.
,
Drenth
J.P.
,
Politei
J.M.
,
Waxman
S.G.
et al
(
2017
)
Familial gain-of-function Nav1.9 mutation in a painful channelopathy
.
J. Neurol. Neurosurg. Psychiatry
88
,
233
240
[PubMed]
148
Estacion
M.
,
Han
C.
,
Choi
J.S.
,
Hoeijmakers
J.G.
,
Lauria
G.
,
Drenth
J.P.
et al
(
2011
)
Intra- and interfamily phenotypic diversity in pain syndromes associated with a gain-of-function variant of NaV1.7
.
Mol. Pain
7
,
92
[PubMed]
149
Dib-Hajj
S.D.
,
Yang
Y.
,
Black
J.A.
and
Waxman
S.G.
(
2013
)
The NaV1.7 sodium channel: from molecule to man
.
Nat. Rev. Neurosci.
14
,
49
62
[PubMed]
150
Hong
S.
,
Morrow
T.J.
,
Paulson
P.E.
,
Isom
L.L.
and
Wiley
J.W.
(
2004
)
Early painful diabetic neuropathy is associated with differential changes in tetrodotoxin-sensitive and -resistant sodium channels in dorsal root ganglion neurons in the rat
.
J. Biol. Chem.
279
,
29341
29350
[PubMed]
151
Black
J.A.
,
Liu
S.
,
Tanaka
M.
,
Cummins
T.R.
and
Waxman
S.G.
(
2004
)
Changes in the expression of tetrodotoxin-sensitive sodium channels within dorsal root ganglia neurons in inflammatory pain
.
Pain
108
,
237
247
[PubMed]
152
Shao
D.
,
Okuse
K.
and
Djamgoz
M.B.
(
2009
)
Protein-protein interactions involving voltage-gated sodium channels: post-translational regulation, intracellular trafficking and functional expression
.
Int. J. Biochem. Cell Biol.
41
,
1471
1481
[PubMed]
153
Leterrier
C.
,
Brachet
A.
,
Fache
M.P.
and
Dargent
B.
(
2010
)
Voltage-gated sodium channel organization in neurons: protein interactions and trafficking pathways
.
Neurosci. Lett.
486
,
92
100
[PubMed]
154
Malhotra
J.D.
,
Kazen-Gillespie
K.
,
Hortsch
M.
and
Isom
L.L.
(
2000
)
Sodium channel subunits mediate homophilic cell adhesion and recruit ankyrin to points of cell-cell contact
.
J. Biol. Chem.
275
,
11383
11388
[PubMed]
155
Lopez-Santiago
L.F.
,
Pertin
M.
,
Morisod
X.
,
Chen
C.
,
Hong
S.
,
Wiley
J.
et al
(
2006
)
Sodium channel β2 subunits regulate tetrodotoxin-sensitive sodium channels in small dorsal root ganglion neurons and modulate the response to pain
.
J. Neurosci.
26
,
7984
7994
[PubMed]
156
Laedermann
C.J.
,
Cachemaille
M.
,
Kirschmann
G.
,
Pertin
M.
,
Gosselin
R.D.
,
Chang
I.
et al
(
2013
)
Dysregulation of voltage-gated sodium channels by ubiquitin ligase NEDD4-2 in neuropathic pain
.
J. Clin. Invest.
123
,
3002
3013
[PubMed]
157
Laedermann
C.J.
,
Decosterd
I.
and
Abriel
H.
(
2014
)
Ubiquitylation of voltage-gated sodium channels
.
Handb. Exp. Pharmacol.
221
,
231
250
[PubMed]
158
Benson
M.D.
,
Li
Q.J.
,
Kieckhafer
K.
,
Dudek
D.
,
Whorton
M.R.
,
Sunahara
R.K.
et al
(
2007
)
SUMO modification regulates inactivation of the voltage-gated potassium channel Kv1.5
.
Proc. Natl Acad. Sci. U.S.A.
104
,
1805
1810
159
Plant
L.D.
,
Dowdell
E.J.
,
Dementieva
I.S.
,
Marks
J.D.
and
Goldstein
S.A.
(
2011
)
SUMO modification of cell surface Kv2.1 potassium channels regulates the activity of rat hippocampal neurons
.
J. Gen. Physiol.
137
,
441
454
[PubMed]
160
Plant
L.D.
,
Marks
J.D.
and
Goldstein
S.A.
(
2016
)
SUMOylation of NaV1.2 channels mediates the early response to acute hypoxia in central neurons
.
Elife
5
,
e20054
[PubMed]
161
Anckar
J.
and
Sistonen
L.
(
2007
)
SUMO: getting it on
.
Biochem. Soc. Trans.
35
,
1409
1413
[PubMed]
162
Wilkinson
K.A.
,
Nishimune
A.
and
Henley
J.M.
(
2008
)
Analysis of SUMO-1 modification of neuronal proteins containing consensus SUMOylation motifs
.
Neurosci. Lett.
436
,
239
244
[PubMed]
163
Wilkinson
K.A.
and
Henley
J.M.
(
2010
)
Mechanisms, regulation and consequences of protein SUMOylation
.
Biochem. J.
428
,
133
145
[PubMed]
164
Mukhopadhyay
D.
and
Dasso
M.
(
2007
)
Modification in reverse: the SUMO proteases
.
Trends Biochem. Sci.
32
,
286
295
[PubMed]
165
Takahashi
Y.
,
Hasegawa-Moriyama
M.
,
Sakurai
T.
and
Inada
E.
(
2011
)
The macrophage-mediated effects of the peroxisome proliferator-activated receptor-gamma agonist rosiglitazone attenuate tactile allodynia in the early phase of neuropathic pain development
.
Anesth. Analg.
113
,
398
404
[PubMed]
166
Meinecke
I.
,
Cinski
A.
,
Baier
A.
,
Peters
M.A.
,
Dankbar
B.
,
Wille
A.
et al
(
2007
)
Modification of nuclear PML protein by SUMO-1 regulates Fas-induced apoptosis in rheumatoid arthritis synovial fibroblasts
.
Proc. Natl Acad. Sci. U.S.A.
104
,
5073
5078
167
Benson
M.
,
Iniguez-Lluhi
J.A.
and
Martens
J.
(
2017
)
SUMO modifications of ion channels
.
Adv. Exp. Med. Biol.
963
,
127
141
[PubMed]
168
Majava
V.
,
Loytynoja
N.
,
Chen
W.Q.
,
Lubec
G.
and
Kursula
P.
(
2008
)
Crystal and solution structure, stability and posttranslational modifications of collapsin response mediator protein 2
.
FEBS J.
275
,
4583
4596
[PubMed]
169
Dustrude
E.T.
,
Moutal
A.
,
Yang
X.
,
Wang
Y.
,
Khanna
M.
and
Khanna
R.
(
2016
)
Hierarchical CRMP2 posttranslational modifications control NaV1.7 function
.
Proc. Natl Acad. Sci. U.S.A.
113
,
E8443
E8452
170
Santolini
E.
,
Puri
C.
,
Salcini
A.E.
,
Gagliani
M.C.
,
Pelicci
P.G.
,
Tacchetti
C.
et al
(
2000
)
Numb is an endocytic protein
.
J. Cell Biol.
151
,
1345
1352
[PubMed]
171
Nishimura
T.
,
Fukata
Y.
,
Kato
K.
,
Yamaguchi
T.
,
Matsuura
Y.
,
Kamiguchi
H.
et al
(
2003
)
CRMP-2 regulates polarized Numb-mediated endocytosis for axon growth
.
Nat. Cell Biol.
5
,
819
826
[PubMed]
172
Woelk
T.
,
Oldrini
B.
,
Maspero
E.
,
Confalonieri
S.
,
Cavallaro
E.
,
Di Fiore
P.P.
et al
(
2006
)
Molecular mechanisms of coupled monoubiquitination
.
Nat. Cell Biol.
8
,
1246
1254
[PubMed]
173
Horvath
C.A.
,
Vanden Broeck
D.
,
Boulet
G.A.
,
Bogers
J.
and
De Wolf
M.J.
(
2007
)
Epsin: inducing membrane curvature
.
Int. J. Biochem. Cell Biol.
39
,
1765
1770
[PubMed]
This is an open access article published by Portland Press Limited on behalf of the Biochemical Society and distributed under the Creative Commons Attribution License 4.0 (CC BY).