Next Article in Journal
Metabolite Profiling of Alangium salviifolium Bark Using Advanced LC/MS and GC/Q-TOFTechnology
Next Article in Special Issue
ATF3-Expressing Large-Diameter Sensory Afferents at Acute Stage as Bio-Signatures of Persistent Pain Associated with Lumbar Radiculopathy
Previous Article in Journal
Erratum: Gruber, E.S.; et al. The Oncogene AF1Q is Associated with WNT and STAT Signaling and Offers a Novel Independent Prognostic Marker in Patients with Resectable Esophageal Cancer. Cells 2019, 8, 1357
Previous Article in Special Issue
pH Mapping of Skeletal Muscle by Chemical Exchange Saturation Transfer (CEST) Imaging
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Peripheral Neuropathic Pain: From Experimental Models to Potential Therapeutic Targets in Dorsal Root Ganglion Neurons

1
Department of Anatomy and Cell Biology, College of Medicine, National Taiwan University, Taipei 10051, Taiwan
2
Department of Life Science, College of Life Science, National Taiwan University, Taipei 10617, Taiwan
3
Department of Anatomy, School of Medicine, College of Medicine, Kaohsiung Medical University, Kaohsiung 80708, Taiwan
4
School of Post-Baccalaureate Medicine, College of Medicine, Kaohsiung Medical University, Kaohsiung 80708, Taiwan
5
Department of Medical Research, Kaohsiung Medical University Hostpital, Kaohsiung 80708, Taiwan
6
Department of Anatomy, School of Medicine, Chung Shan Medical University, Taichung 40201, Taiwan
7
Department of Medical Education, Chung Shan Medical University Hospital, Taichung 40201, Taiwan
8
Akouos, Inc., Boston, MA 02210, USA
9
Department of Neurology, National Taiwan University Hospital, Taipei 10002, Taiwan
10
Graduate Institute of Brian and Mind Sciences, College of Medicine, National Taiwan University, Taipei 10051, Taiwan
11
Center of Precision Medicine, College of Medicine, National Taiwan University, Taipei 10055, Taiwan
*
Author to whom correspondence should be addressed.
Cells 2020, 9(12), 2725; https://doi.org/10.3390/cells9122725
Submission received: 14 October 2020 / Accepted: 16 December 2020 / Published: 21 December 2020

Abstract

:
Neuropathic pain exerts a global burden caused by the lesions in the somatosensory nerve system, including the central and peripheral nervous systems. The mechanisms of nerve injury-induced neuropathic pain involve multiple mechanisms, various signaling pathways, and molecules. Currently, poor efficacy is the major limitation of medications for treating neuropathic pain. Thus, understanding the detailed molecular mechanisms should shed light on the development of new therapeutic strategies for neuropathic pain. Several well-established in vivo pain models were used to investigate the detail mechanisms of peripheral neuropathic pain. Molecular mediators of pain are regulated differentially in various forms of neuropathic pain models; these regulators include purinergic receptors, transient receptor potential receptor channels, and voltage-gated sodium and calcium channels. Meanwhile, post-translational modification and transcriptional regulation are also altered in these pain models and have been reported to mediate several pain related molecules. In this review, we focus on molecular mechanisms and mediators of neuropathic pain with their corresponding transcriptional regulation and post-translational modification underlying peripheral sensitization in the dorsal root ganglia. Taken together, these molecular mediators and their modification and regulations provide excellent targets for neuropathic pain treatment.

1. Introduction

1.1. Overview of Neuropathic Pain

The prevalence of chronic pain with neuropathic characteristics has been reported to be over 7% of the general population in the world but this number could be underestimated because of the difficulty in characterization and poorly understood mechanisms [1]. The International Association for the Study of Pain (IASP) indicated that health-related quality of life is low in people suffering from neuropathic pain. Although scientists have been working hard to investigate relevant signal transductions in neuropathic pain for decades, only <50% patients could have a relief from neuropathic pain after medications due to the complicated signal transductions playing in neuropathic pain pathways and circuits [2,3]. The lesions in the central or peripheral nervous system reduce the threshold and increase the responses to the nociceptive stimulation. In peripheral nerve injury, the nociception is conveyed by the primary sensory neurons in the dorsal root ganglia (DRG) and back to secondary sensory neurons.
Nerve fibers are classified into different types including A, B, and C fibers according to their diameter, myelination status, and conduction velocity. In addition to the lesions in the central nervous system causing neuropathic pain, peripheral nerve degeneration (peripheral neuropathy) also induces neuropathic pain by affecting small myelinated Aδ fibers and unmyelinated C fibers [4]. The small myelinated Aδ fibers and unmyelinated C nerve fibers are mainly affected in peripheral neuropathic pain syndromes. Clinical symptoms of peripheral neuropathy include tingling, burning, and numbness due to dysregulation of nerve fibers [5]. Till now, although antidepressants, anticonvulsants, and opioids were recommended in the guidelines for treating neuropathic pain, the responses of treatments are not satisfactory and large scale of studies are continuously conducted to explore the whole pictures of neuropathic pain in different fields.
To investigate the detailed mechanisms of neuropathic pain, well-established in vivo models have been applied to mimic different etiologies of neuropathic pain such as the injury-induced and systemic exposure to neurotoxins. In spite of the fact that the proportion of peripheral sensitization driven neuropathic pain is difficult to distinguish, focusing on peripheral desensitization could be a phased strategy for treating peripheral neuropathic pain before intervene into central sensitization. Compared to the central nerve system, the permeability in DRG is relatively higher due to the incomplete blood–nerve barrier [6,7]. Neurotoxins could easily penetrate the blood–nerve barrier into the DRG and cause neuronal excitability [8]. The level of molecular expression in DRG neurons at the levels of gene and protein vary in neuropathic pain [9,10,11]. Converging researches showed that manipulation of DRG could relieve pain through changing the excitability and inhibiting signaling pathways [12,13,14,15,16,17]. Furthermore, the therapeutic region could be focal and easily restricted by targeting specific levels of DRG. Therefore, the molecules in DRG could be potential therapeutic targets in neuropathic pain. In this review, we will summarize the pain related molecules in neuropathic pain and focus on the peripheral sensitization within post-translational control and transcriptional modifications in DRG.

1.2. Etiology and Patholoigcal Approches

The challenge in managing neuropathic pain is to identify the exact etiology for appropriate treatment due to different etiologies affect various types of never fibers [18]. Both large and small fibers in the peripheral nervous system are affected in injury-induced neuropathic pain models. However, some disease such as chemotherapy neuropathy, vasculitic neuropathy, and Charcot–Marie–Tooth disease mainly target large diameter nerve fibers. Resiniferatoxin (RTX), one of the neurotoxins which belongs to capsaicin analogues that induces the desensitization in nociceptive neuron. However, RTX only affects small diameter nerve fibers specifically. Intraepidermal nerve fiber (IENF) density from the skin biopsy with immunohistochemistry staining of protein gene product 9.5 could be a standard and useful tool to estimate the degeneration of small-diameter sensory neurons in these neuropathic pain models [19]. Whereas, sural sensory nerve action potential (SNAP) and sural nerve morphometry could be performed to examine the pathological finding in larger fiber neuropathy [20].

2. Neuropathic Pain Models

2.1. Injury Induced

Several surgical neuropathic pain models have been developed decades and are widely used to mimic injury-induced neuropathic pain, including spinal nerve ligation (SNL), chronic constriction injury (CCI), partial sciatic nerve injury, and spared nerve injury (SNI) [21,22,23]. Each model has stable, long-lasting, and reproducible characteristics of neuropathic pain behavior, such as mechanical allodynia and thermal hyperalgesia but with corresponding unique pain behavior patterns [24]. The major factor distinguishing various injury-induced neuropathic pain models resides in the location of manipulations leading to similarities and differences (Table 1). We first will focus on the application of these injury induced neuropathic pain models.
Decompression is usually performed to relief the compression-induced neuropathic pain clinically. Although the variations exist in each operation, CCI produces spontaneous pain characteristic with four loose ligatures on the sciatic nerve. We performed the decompression at postoperative week four by removing the ligatures to investigate the mechanisms and temporal course of neuropathic pain [25]. Plasticity in the dorsal horn and partial reinnervation of the skin contributed to the normalization of neuropathic pain behavior [25,26].
Compared to CCI, the surgical procedure of SNL is much more complex and time consuming [24]. However, the variation in SNL is much lower given that same degree of injured spinal nerves in each operation. SNL separates the injured and uninjured DRG by ligating spinal nerve at the specific level and provides an opportunity to inject chemicals or gene constructs through the spinal nerve [27]. By combining spinal nerve injection and SNL model, the efficiency of therapeutic agents on injured DRG can be easily addressed.

2.2. Systemic Exposure Neuropathic Pain Model

In about 50% of sensory neuropathy, small fiber is predominantly affected [28]. Various diseases affect specific type of nerve fibers. For example, taxanes and vincristine mainly affect large fiber, whereas small fibers are affected at the early stage of diabetes mellitus [29,30]. The surgical neuropathic pain models that are mentioned above usually damage the nerve fiber without fiber type preference. Therefore, to establish suitable neuropathic pain model that mimics small or large fiber neuropathy is necessary for studying the disease like chemotherapy-induced neuropathy and painful diabetic peripheral neuropathy.

2.2.1. Resiniferatoxin

Resiniferatoxin, an analog of capsaicin, is widely used for generating a small fiber neuropathy in rodents [31]. Hsieh et al. established a small fiber neuropathic pain model by using a single dosage of RTX. Experimental mice developed a pure small fiber neuropathy one week after an exposure to 50 μg/kg of RTX through intraperitoneal injection [32]. The number of injury marker (activating transcription factors 3 (ATF3)) was significantly increased only in small DRG neurons and the large DRG neurons remained ATF3 (−). Reduced of IENF density, mechanical allodynia, and thermal hypoalgesia were the characteristic pathology and behaviors in this model (Figure 1). To investigate the strategy for small fiber reinnervation in RTX-induced small fiber neuropathy, 4-methylcatechol (4-MC) could be used as a positive control for screening therapeutic candidate chemicals [31,33].

2.2.2. Acrylamide

Acrylamide is one of the neurotoxins that targets the central and peripheral nervous system [34]. Both large and small nerve fibers could be affected by acrylamide through different exposure routes [35,36]. The unique pathology of acrylamide-induced neuropathy (“dying-back” phenomena) indicated the early degeneration of axons at the distal end. Each stage of neuropathy induced by acrylamide is distinct and could be classified by the neurological symptoms [36]. With accumulated acrylamide (35–140 mg/kg) by appropriate low-dose and repeated exposure, experimental animals showed mechanical allodynia and thermal hyperalgesia but without motor deficits [37]. Induction of massive oxidative stress is one of the neurotoxic mechanisms [38]. Treating antioxidants could reduce the accumulation of reactive oxygen species (ROS) in in vitro cell models and reverse the neurotoxicity that induced by acrylamide in in vivo animal model [39,40].

2.2.3. Chemotherapy-Induced Peripheral Neuropathy (CIPN)

Various adverse effects occur when patients receive chemotherapy. About 60% of patients on chemotherapy in one month develop a chemotherapy-induced peripheral neuropathy with the “glove and stocking” type distribution [41]. The symptoms of CIPN sometimes might continue for several months even after the chemotherapy is completed. Different chemotherapy drugs target different types of nerve fibers [42]. The platinum-based drugs affect the large fibers and vincristine affects the nerve fibers without fiber type preference. Not only the type of fibers but also the target site in the neurons could be used to distinguish these chemotherapy drugs. Taxanes and vincristine target axons, whereas platinum-based drugs (cisplatin and oxaliplatin) target the soma of DRG [43,44,45]. The blood–nerve barrier in DRG is relatively permeable which leads to the penetrating of chemotherapeutic agents from the blood vessel into DRG [6]. Axonal demyelination and degeneration accompanied with mitochondrial dysfunction, axon transport inhibition, and calcium homeostasis are the major pathological findings in CIPN [46]. To date, there are no effective strategies to prevent CIPN during chemotherapy. Therefore, to establish a neuropathic pain model mimic CIPN is important for investigating the detailed mechanisms of CIPN.
The cumulative doses of these chemotherapy drugs (paclitaxel: 4 mg/kg, cisplatin: 23 mg/kg, and vincristine: 200 μg/kg) induced CIPN in animals [47]. Since the demyelinated large fibers are observed in CIPN, a semi-thin section for nerve morphometric analysis could be used for monitoring the status of demyelination in large fiber by calculating the number and the diameter of myelinated and demyelinated nerve fibers [48]. Combining CIPN models with electrophysiological examinations and nerve morphometry, Chine et al. observed that the expression of a regeneration-related protein, heat shock protein 27, in neurons which could reverse the symptoms of paclitaxel-induced CIPN by inhibiting the apoptosis, demyelination, and mitochondria swelling [49].

3. Molecular Targets in Dorsal Root Ganglia

Numerous molecules and pathways have been indicated as potential biomarkers or regulators in neuropathic pain [50]. This review aims to focus on the differences in the expression levels of well-established mediators in the DRG. These renowned mediators include P2X receptor family, the transient receptor potential (TRP) channels, voltage-gated sodium channels (NaVs), and voltage-gated calcium channels (CaVs) (Table 2 and Table 3).

3.1. Purinergic P2X Receptors (P2Xs)

The P2X family receptors belong to cation-permeable ionotropic receptor that can be activated by ATP. P2X receptors are expressed in the spinal cord, dorsal horn, and DRG neurons [51,52,53]. Accumulating evidence has indicated that P2X receptors, in particular, homomeric purinergic receptor P2X3 (P2X3) and heteromeric P2X2/3 receptors, are involved in spinal nociceptive transmission [54]. Thus, P2X2/3 have been proposed as potential therapeutic targets for peripheral neuropathic pain.
The expression level of P2X receptors is variable in different models of neuropathic pain. Rats with SNL had upregulated, downregulated, or unchanged levels of P2X3 expression in DRG [55,56,57]. Nevertheless, researchers have demonstrated that P2X3 exhibits an upregulation in the small to medium size neurons of lumbar DRGs in CCI-treated rats [51,52,53,58]. In the SNI model, the expression level of P2X3 was unchanged in rats. The reason behinds these differences in neuropathic pain models have not yet been fully understood. Additional investigations are required to investigate the roles of P2X receptors in injury-induced neuropathic pain models.
Hsieh et al. had shown that upon RTX exposure, peripherin (+) neurons were reduced and P2X3 (+) neurons were increased [59]. The upregulation of P2X3 (+) neurons was correlated with mechanical allodynia. These findings suggest that P2X3 is responsible for peripheral sensitization of neuropathic pain. Hori et al. found that cisplatin injection upregulated P2X3 expression [60]. This study provided findings of the involvement of P2X3 receptor in neuropathic pain.

3.1.1. Transient Receptor Potential Receptors

Transient receptor potential receptors are non-selective channels that can induce painful sensations by detecting physical and chemical stimuli in the environment. In the primary afferent neurons of DRG, these receptors act as transducers for various kinds of stimuli including mechanical, thermal, and chemical stimuli. TRP receptors are classified into different subfamilies, and each subfamily is slightly different, resulting in detecting the diverse sensory stimuli. Of all of the subfamilies, the transient receptor potential vanilloid (TRPV), transient receptor potential ankyrin (TRPA), and transient receptor potential melastatin (TRPM) contribute to neuropathic pain [61]. In addition to being activated by their respective agonists, different TRP receptors respond to their unique temperature thresholds. For example, TRPV subtype 1 (TRPV1) channels are activated by elevating the temperature above 43 °C. Due to this special property, TRPV1 has been associated with thermal hyperalgesia in animal models [62,63]. The TRPA subtype 1 (TRPA1) channel, on the other hand, has a temperature threshold of 18 °C or under and thus is associated with cold hyperalgesia [63]. TRPM subtype 8 (TRPM8) channels are also temperature-gated with the temperature threshold of 23–28 °C. In injury-induced neuropathic pain models, the expression levels of TRPV1 were upregulated in SNL and CCI but downregulated in SNI [56,64,65,66]. Due to their roles in neuropathic pain, TRP channels have become one of the therapeutic targets. For one, by inhibiting calmodlin-dependent protein kinass II (CAMKII) expression and extracellular signal regulated protein kinase 2 (ERK2) phosphorylation, researchers had successfully silenced the TRPV1 channels and thus attenuated neuropathic pain in rats [67]. Consequently, understanding the altered expression of TRP channels in various neuropathic pain models is essential in developing successful therapy.

3.1.2. Voltage-Gated Sodium Channels (NaVs)

Typically, there are nine genes in the mammals that encode the different subunits of the voltage-gated sodium channels, NaV1.1–NaV1.9. These NaVs can be further categorized into two groups, tetrodotoxin-sensitive (TTX-S) and tetrodotoxin-resistant (TTX-R) based on their respective sensitivity to the neurotoxin tetrodotoxin (TTX). DRG neurons have the most diverse range of NaV subtypes compared to other neuronal cell types, up to five subtypes [68,69]. Matured DRG neurons have the capability to express both TTX-S (NaV1.1, NaV1.6, and NaV1.7) and TTX-R (NaV1.8 and NaV1.9) subtypes of NaVs. Among these NaVs, NaV1.7, NaV1.8, and NaV1.9 are specific subtypes in the neurons of the peripheral nervous system.
These NaVs have been linked with neuropathic pain. Several neuropathic pain models have shown to alter the expressions of NaVs at the transcription level i.e., the transcription can be “turned on” or “turned off” in an event of injury [70]. Increased expression of NaVs have been linked with painful diabetic neuropathy and injury-induced neuropathy [71,72,73]. Thus, the alternation in the expression of NaVs following neuropathic pain models is crucial in the understanding of neuropathic pain.

3.1.3. Voltage-Gated Calcium Channels

Voltage-gated calcium channels are grouped into two classes: High-voltage activated type, the L, N, P/Q, and R type, and low-voltage activated type, also known as T-type [74]. All CaVs, except T-type CaVs, consist of three subunits [75], the α1 channel-forming subunit, the intracellular β subunit, and the α2δ subunit that consists of two disulfide-linked polypeptides (α2 and δ) encoded by the same gene [76]. These subunits may play an important role in neuropathic pain.
T-type CaVs, namely CaV3.2 and CaV3.3, are expressed in small and medium size isolectin B4 (IB4) (+) DRG neurons [77,78,79]. CaV3.2 plays an important role in antinociceptive, antihyperalgesic, and antiallodynic behavior after neuropathic pain of CCI [80], providing evidence that T-type CaVs can be potential therapeutic targets of neuropathic pain.
Both mRNA and protein levels of the α2δ-1 subunit of voltage-gated calcium channels were upregulated in SNL rats. Furthermore, the elevation of α2δ-1 was correlated with mechanical allodynia [81]. These results suggest that the α2δ-1 subunit could be a therapeutic target for peripheral neuropathic pain. However, in CCI-treated mice, several species of the mRNA that encode the α1 subunit of CaVs were downregulated [82]. The fluctuations of these subunits are considered to modulate the functions of CaVs. In the SNI model, T-type CaV3.2 receptors were upregulated [80]. The post-injury alteration in α2δ level, combined with the successful prevention of injury-induced pain by intrathecal infusion of gabapentin, provides convincing evidence for therapeutic possibilities of CaVs.

4. Epigenetic, Transcriptional, and Post-Translational Modification in Neuropathic Pain

In addition to the protein dogma theory, there are reports highlighting the participating epigenetic changes, translational modification, and post-translational control in various biological processes such as development, differentiation, proliferation, cell death, and cancer progression [83,84]. Several lines of evidence indicated that these regulations also participate in neuropathic pain [85,86,87]. Hence, personalize medicine could provide the optimized medical decision for different patients according to their personal genetic background such as single nucleotide polymorphism [88]. In this part, we will focus on several topics in peripheral neuropathic pain; (1) DNA methylation and acetylation, (2) the interactions between phosphorylated extracellular signal-regulated kinase (p-ERK) and transcriptional factors (TFs), and (3) ubiquitination and deubiquitination.

4.1. DNA Methyltransferase (DNAMT) and Histone Deacetylase (HDAC)

Epigenetic changes including DNA methylation, and histone acetylation regulate the gene expression without changing the coding sequence. The basic unit of chromatin, nucleosome, is composed of a segment of DNA and two copies of four histone proteins (H2A, H2B, H3, and H4) [89]. The folding pattern of chromatin DNA governs to the gene transcription efficiency. Chromatin enriched with methylation likely reduces the transcription efficiency. In contrast, the acetylated chromatin with less condensed DNA structures presents the most active transcriptional region [90]. Once the homeostasis of DNA methylation and acetylation is disrupted, the dysregulated gene expression will lead to various pathology or diseases such as cancer.

4.1.1. DNA Methylation and Demethylation

Accumulated evidence has showed the potential contribution of DNA methylation and demethylation to the regulation of neuropathic pain behaviors [85,91]. DNA methylation was decreased in DRG after SNL and the hypomethylation leading to mechanical hypersensitivity by (1) intrathecal injection of DNA methyltransferase inhibitor, RG108 or (2) giving a methyl donor-deficient diet. The expression of genes related to voltage-gated and ligand-gated ion channels in DRG were identified and decreased based on the comparison between the RNA-seq data and the DNA methylation data [92]. Zhang et al. demonstrated that the demethylation on the promoter region of the P2X3 receptor was increased four weeks after injected a single dosage of streptozocin (STZ). The demethylated region of the P2X3 receptor could increase nuclear factor kappa B (NF-κB) binding to the P2X3 promoter followed with an increase in the expression level of P2X3 receptor [93]. The decreased voltage-dependent potassium channel subunit 2 (Kcna2) in DRG was observed in the neuropathic pain model of SNL and CCI. The expression of Kcna2 could be regulated by octamer transcription factor (OCT1) that binds to the promoter region of DNMT3a on day three post-SNL. The expression level of Kcna2 could be rescued after treating DNA methylation blocker [94].

4.1.2. Histone Deacetylase Inhibitors

Histone deacetylase inhibitors are well developed and frequently used in cancer therapy. In recent years, HDACs have been implicated in the mechanisms of neuron degeneration and regulation of the neuronal plasticity [95,96,97]. For example, HDAC5 was increased in DRG on day seven after partial sciatic nerve ligation. The increased HDAC5 could bind to the Sox10 promoter region [98]. Matsushita et al. demonstrated that the activity of C-fiber could be reversed after treating HDAC inhibitors in partial sciatic nerve ligation-induced neuropathic pain model [99]. Mechanical and thermal hypersensitivities induced by ligation or transection of the spinal nerve were decreased after a treatment with class I HDAC inhibitors through the intrathecal route [100].

4.2. p-ERK and Transcriptional Factors

Mitogen-activated protein kinases (MAPKs) and TFs are molecules that responded rapidly to stress, hypoxia, and injury. Both MAPKs, especially ERK, and various TFs involve in peripheral neuropathic pain. p-ERK up-regulation in the dorsal horn neurons of the spinal cord contribute to the neuropathic pain of different nerve injuries [26,101,102]. The pain behaviors induced by chronic compression of the spinal nerve was highly correlated with an increase in phosphorylated-MAPKs including p-ERK, phosphorylated c-Jun N-terminal kinase (p-JNK), and phosphorylated p38 MAPK (p-p38)) in DRG at four days after surgery. The pain behaviors were reversed after treating MAPKs inhibitors [103]. Specific phosphorylation on the kinase insert domain in the nuclear translocation sequence, p-ERK interacts with a nuclear importing protein (importin-7) and then translocates from the cytosol into the nucleus [104]. Nuclear p-ERK regulates transcription factors and regulatory proteins to influence neuropathic pain mediators such as proto-oncogene c-fos and cAMP-response element binding protein (CREB) [105,106]. Upon peripheral nerve injury, these TFs were upregulated and translocated from the cytosol into the nucleus. The activated nuclear c-fos and CREB is regulated by p-ERK and then mediates the downstream gene involve in neuropathic pain [107].
p-ERK also regulates another transcription factor, activating transcription factors 3 (ATF3), a key mediator in nerve injury and regeneration [108,109]. Our previous study revealed that the increased co-localization of ATF3 and P2X3 on day seven after RTX treatment was highly correlated with resiniferatoxin induced neuropathic pain behaviors [59]. Ding et al. also demonstrated that ATF3 was increased three weeks after operation and participated in P2X3-induced endometriosis pain through ATF3/AP-1 activation [110]. Therefore, phosphorylation of ERK could be a potential therapeutic target for treating neuropathic pain in the future.

4.3. Ubiquitination and Deubiquitination

The ubiquitination and deubiquitination keep a delicate balance for post-translational modification system that facilitates the proteasomal pathway to regulate intracellular protein levels. Both ubiquitination and deubiquitination involve in neuropathic pain. E3 ubiquitin ligase regulates ion channel proteins [111,112,113]. For example, Nedd4-2 (neural precursor cell expressed, developmentally down-regulated 4-like, E3 ubiquitin protein ligase), an E3 ubiquitin ligase, is decreased on day 7 after SNI. The decreased Nedd4-2 is responsible for ubiquitination of NaVs, and caused dysregulation of NaV1.7 and NaV1.8 in the SNI model [73]. García–Caballero et al. (2014) showed that ubiquitin-specific protease 5 (USP5), a deubiquitinating enzyme, is increased on day 14 after CCI and modulates T-type CaV3.2 [114]. The depletion of USP5 in DRG successfully alleviated pain hypersensitivity in SNI mice, implicating the causal relationship between USP5 and pain induced by neuropathic models.

5. Current Challenges to Manipulate on DRG in In Vivo Models

DRGs which contain pseudounipolar neurons that perceive sensations from the terminal ending of nerve fiber then transduce back to the spinal cord [115]. After receiving the stimuli, DRG neurons will be activated by increasing the neuronal excitability and nociceptive signals [116]. Thus, to manipulate on DRG is a promising approach for investigating the molecular mechanisms of neuropathic pain. Previous studies revealed that direct injection into the DRG might trigger inflammatory response, while the intrathecal route might cause the injected component diffusible in multiple levels of the spinal cord [117,118,119]. We have established the spinal nerve injection technique to delivery gene constructs at a single site without removing the vertebrae and hence provided a feasible approach with a low risk of inflammation and injury [27]. There was no activation of ATF3 and ionized calcium binding adaptor molecule 1 (Iba1) in DRG neurons after spinal nerve injection in comparison with intra-DRG injection. Although the technique of spinal nerve injection presents much more challenge compared to intrathecal injection, spinal nerve injection is suitable for delivering drugs or gene construct into DRG in in vivo studies because some etiologies causing peripheral neuropathic pain is local but not systemic.
Gene therapy is a promising therapy for treating neuropathic pain. Several AAV clinical trials were already undergoing for Alzheimer’s disease, Parkinson’s disease, rheumatoid arthritis, etc. [120]. The serotype of AAV most widely used in clinical trial is AAV2 that work well in the CNS but with poor infection efficiency in DRG [121,122]. Several serotypes of AAV affect the infection efficiency, and the promoters on the construct also show huge variations on the ability of driving gene expression. Tissue specific promoters are designed and cloned into construct backbone to deliver gene expression to specific cell type in tissues [123]. Exposure route is another issue for concern in gene therapy. High titer of AAV virus particles must be used in systemic injection in which severe immune response might be triggered [124]. Because multiple systems or tissues are affected in many diseases, local injection the AAV virus particles might only be feasible in certain diseases such as gene mutation-related congenital blindness [125]. The above-mentioned issues might be considered carefully for selecting suitable AAV backbone, exposure rout, and serotype on clinical applications in the future.

6. Conclusions

The guidelines for treating neuropathic pain have been published by several international and regional professional associations. Antidepressants, anticonvulsants, and opioids are commonly used for treating neuropathic pain. However, these current therapy drugs like opioids do not directly target on pain-related molecules. The most challenges of the treatment on neuropathic pain are adverse effects, a lack of specificity, addiction, and inefficiency of pain relief. Although several voltage-gated sodium channel blockers were used to treat neuropathic pain, but the outcomes of these clinical trials still face the issue mentioned above [126]. Dexpramipexole, a specific blocker (NaV1.7 and NaV1.8) that effectively relieves neuropathic pain in various neuropathic pain models with well tolerance in several clinical trials, is undergoing clinical trial [127,128,129,130]. Therefore, targets on voltage-gated sodium channels could be the potential therapeutic targets in neuropathic pain. Due to the DRG is the first gate that connects the spinal cord and terminal nerve ending and receives the stimuli from nerve injury. Investigating the functions of pain-related genes and proteins in DRG through these different neuropathic pain models is necessary for screening potential targets. High throughput screening such as next generation sequencing, proteomics, and metabolomics are well-established approaches for searching novel candidate therapeutic targets. Combination of current therapeutic drugs with targeting novel molecules at the same time will reduce the side effects and be an alternative therapy in the future.
Table 2. Molecular mediators of neuropathic pain in focal neuropathy models.
Table 2. Molecular mediators of neuropathic pain in focal neuropathy models.
DRGP2XsTRPsNaVsCaVs
Injury-Induced
SNLP2X3 ↑ (IR: d7, mRNA: d1) [55,56,57]
P2X3 ↓ (IR: d14, mRNA: d7) [55,57]
P2X3—(IR: d3, mRNA: d14) [56]
P2X4—(mRNA: d1, d7) [57]
P2X5 ↑ (mRNA: d1, d7) [57]
P2X6 ↓ (mRNA: d7) [57]
P2X6—(mRNA: d1) [57]
TRPV1 ↑ (IR: d3, mRNA: d1-28) [56]
TRPA1 ↑ (IR: d11, mRNA) [131]
NaV1.7 ↓ (L5, IR: d7, mRNA: d7) [132]
NaV1.7 ↑ (L4, IR: d7, mRNA: d7) [132]
NaV1.3 ↑ [133]
NaV1.1 ↓ [133]
NaV1.6 ↓ [133]
NaV1.7 ↓ [133]
NaV1.8 ↓ [133]
NaV1.9 ↓ [133]
α2δ-1 subunit ↑ (mRNA: d2) [81]
CCIP2X2—(IR: d18) [134]
P2X2 ↑ (mRNA: d14) [135]
P2X3 ↑ (IR: d14, d18, mRNA: d14) [51,52,53,58,135]
P2X4—(IR: d18) [134]
P2X4 ↓ (IR:d18, mRNA: d14) [51,53]
P2X5—(mRNA: d14) [134]
P2X6—(IR: d18) [134]
P2X6 ↑ (IR: d18) [53]
TRPM8 ↑ (IR: d4, d7, d10, d14, mRNA: d14) [64,136]
TPRMP—(mRNA: d7) [136]
TRPA1 ↑ (mRNA: d7, d14,
d16) [136,137]
TRPV1 ↑ (mRNA: d16) [137]
TRPV1—(mRNA: d7, d14) [136]
NaV1.7 ↑ (IR: d7,
mRNA: d7) [132]
α 1 ↓ (mRNA: d7) [82]
SNIP2X3—(IR: d7) [135]TRPM8 ↓ [138]
TRPV1 ↓ [138]
TRPA1 ↓ [138]
Rats (protein) [139]
NaV1.3 ↑
NaV1.7 ↓
NaV1.8 ↓
NaV1.9 ↓
Mice (mRNA) [140]
NaV1.1 ↓
NaV1.2 −
NaV1.3 ↓
NaV1.6 ↓
NaV1.7 ↓
NaV1.8 ↓
NaV1.9 ↓
Cav3.2 ↑ [80]
Abbreviation: IR, immunoreactivity; mRNA, messenger RNA. Notations: ↑, upregulation post-exposure of neuropathic pain. ↓, downregulation post-exposure of neuropathic pain. −, no significant change post-exposure of neuropathic pain. N/A, no direct evidence at the current time.
Table 3. Molecular mediators of neuropathic pain in systemic small fiber neuropathy models.
Table 3. Molecular mediators of neuropathic pain in systemic small fiber neuropathy models.
DRGP2XsTRPsNaVsCaVs
Systemic Small Fiber Neuropathy
RTXP2X3 ↑ (IR: d7) [59]TRPV1 ↓ (mRNA: d3)
[53,59,136]
TRPA1 ↓ (mRNA: d3)
[136,141]
TRPM8—(mRNA: d3)
[136]
N/AN/A
CisplatinP2X3 ↑ (mRNA: d14) [60]TRPV1 ↑ [142]
TRPA1 ↑ [142]
TRPM8—[142]
N/ACaV2.2 ↑ (IR: d3) [143]
Notations: ↑, upregulation post-exposure of neuropathic pain. ↓, downregulation post-exposure of neuropathic pain. —, no significant change post-exposure of neuropathic pain. N/A, no direct evidence at the current time.

Author Contributions

Conceptualization: S.-T.H. Review literature: T.-Y.Y. and I.-W.L. Writing—original draft preparation: T.-Y.Y. and I.-W.L. Writing—review and editing: T.-Y.Y. and S.-T.H. Prepared figure and tables: Y.-L.H., T.-J.T., and H.C. Supervision: S.-T.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

This work was supported by grants from Ministry of Science and Technology (MOST 107-2320-B-002-043-MY3, 107-2314-B-002-069-MY2, 107-3017-F-002-002), Ministry of Education (107L9014-2) and National Taiwan University Hospital (UN109-013).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Van Hecke, O.; Austin, S.K.; Khan, R.A.; Smith, B.H.; Torrance, N. Neuropathic pain in the general population: A systematic review of epidemiological studies. Pain 2014, 155, 654–662. [Google Scholar] [CrossRef]
  2. Colloca, L.; Ludman, T.; Bouhassira, D.; Baron, R.; Dickenson, A.H.; Yarnitsky, D.; Freeman, R.; Truini, A.; Attal, N.; Finnerup, N.B.; et al. Neuropathic pain. Nat. Rev. Dis. Primers 2017, 3, 17002. [Google Scholar] [CrossRef] [Green Version]
  3. Hansson, P.T.; Attal, N.; Baron, R.; Cruccu, G. Toward a definition of pharmacoresistant neuropathic pain. Eur. J. Pain 2009, 13, 439–440. [Google Scholar] [CrossRef]
  4. Omori, S.; Isose, S.; Misawa, S.; Watanabe, K.; Sekiguchi, Y.; Shibuya, K.; Beppu, M.; Amino, H.; Kuwabara, S. Pain-related evoked potentials after intraepidermal electrical stimulation to Adelta and C fibers in patients with neuropathic pain. Neurosci. Res. 2017, 121, 43–48. [Google Scholar] [CrossRef]
  5. Lacomis, D. Small-fiber neuropathy. Muscle Nerve 2002, 26, 173–188. [Google Scholar] [CrossRef]
  6. Abram, S.E.; Yi, J.; Fuchs, A.; Hogan, Q.H. Permeability of injured and intact peripheral nerves and dorsal root ganglia. Anesthesiology 2006, 105, 146–153. [Google Scholar] [CrossRef]
  7. Jacobs, J.M.; Macfarlane, R.M.; Cavanagh, J.B. Vascular leakage in the dorsal root ganglia of the rat, studied with horseradish peroxidase. J. Neurol. Sci. 1976, 29, 95–107. [Google Scholar] [CrossRef]
  8. Sapunar, D.; Ljubkovic, M.; Lirk, P.; McCallum, J.B.; Hogan, Q.H. Distinct membrane effects of spinal nerve ligation on injured and adjacent dorsal root ganglion neurons in rats. Anesthesiology 2005, 103, 360–376. [Google Scholar] [CrossRef]
  9. Chen, C.J.; Liu, D.Z.; Yao, W.F.; Gu, Y.; Huang, F.; Hei, Z.Q.; Li, X. Identification of key genes and pathways associated with neuropathic pain in uninjured dorsal root ganglion by using bioinformatic analysis. J. Pain Res. 2017, 10, 2665–2674. [Google Scholar] [CrossRef] [Green Version]
  10. Maratou, K.; Wallace, V.C.; Hasnie, F.S.; Okuse, K.; Hosseini, R.; Jina, N.; Blackbeard, J.; Pheby, T.; Orengo, C.; Dickenson, A.H.; et al. Comparison of dorsal root ganglion gene expression in rat models of traumatic and HIV-associated neuropathic pain. Eur. J. Pain 2009, 13, 387–398. [Google Scholar] [CrossRef] [Green Version]
  11. Komori, N.; Takemori, N.; Kim, H.K.; Singh, A.; Hwang, S.H.; Foreman, R.D.; Chung, K.; Chung, J.M.; Matsumoto, H. Proteomics study of neuropathic and nonneuropathic dorsal root ganglia: Altered protein regulation following segmental spinal nerve ligation injury. Physiol. Genom. 2007, 29, 215–230. [Google Scholar] [CrossRef] [Green Version]
  12. Puljak, L.; Kojundzic, S.L.; Hogan, Q.H.; Sapunar, D. Targeted delivery of pharmacological agents into rat dorsal root ganglion. J. Neurosci. Methods 2009, 177, 397–402. [Google Scholar] [CrossRef] [Green Version]
  13. Chen, H.P.; Zhou, W.; Kang, L.M.; Yan, H.; Zhang, L.; Xu, B.H.; Cai, W.H. Intrathecal miR-96 inhibits Nav1.3 expression and alleviates neuropathic pain in rat following chronic construction injury. Neurochem. Res. 2014, 39, 76–83. [Google Scholar] [CrossRef]
  14. Xu, X.; Fu, S.; Shi, X.; Liu, R. Microglial BDNF, PI3K, and p-ERK in the Spinal Cord Are Suppressed by Pulsed Radiofrequency on Dorsal Root Ganglion to Ease SNI-Induced Neuropathic Pain in Rats. Pain Res. Manag. 2019, 2019, 5948686. [Google Scholar] [CrossRef]
  15. Wang, Y.; Xue, M.; Xia, Y.; Jiang, Q.; Huang, Z.; Huang, C. Electroacupuncture treatment upregulates alpha7nAChR and inhibits JAK2/STAT3 in dorsal root ganglion of rat with spared nerve injury. J. Pain Res. 2019, 12, 1947–1955. [Google Scholar] [CrossRef] [Green Version]
  16. Liem, L.; Russo, M.; Huygen, F.J.; Van Buyten, J.P.; Smet, I.; Verrills, P.; Cousins, M.; Brooker, C.; Levy, R.; Deer, T.; et al. One-year outcomes of spinal cord stimulation of the dorsal root ganglion in the treatment of chronic neuropathic pain. Neuromodulation 2015, 18, 41–48. [Google Scholar] [CrossRef]
  17. Van Buyten, J.P.; Smet, I.; Liem, L.; Russo, M.; Huygen, F. Stimulation of dorsal root ganglia for the management of complex regional pain syndrome: A prospective case series. Pain Pract. 2015, 15, 208–216. [Google Scholar] [CrossRef] [PubMed]
  18. Chiang, M.C.; Tseng, M.T.; Pan, C.L.; Chao, C.C.; Hsieh, S.T. Progress in the treatment of small fiber peripheral neuropathy. Expert Rev. Neurother. 2015, 15, 305–313. [Google Scholar] [CrossRef]
  19. McCarthy, B.G.; Hsieh, S.T.; Stocks, A.; Hauer, P.; Macko, C.; Cornblath, D.R.; Griffin, J.W.; McArthur, J.C. Cutaneous innervation in sensory neuropathies: Evaluation by skin biopsy. Neurology 1995, 45, 1848–1855. [Google Scholar] [CrossRef]
  20. Herrmann, D.N.; Griffin, J.W.; Hauer, P.; Cornblath, D.R.; McArthur, J.C. Epidermal nerve fiber density and sural nerve morphometry in peripheral neuropathies. Neurology 1999, 53, 1634–1640. [Google Scholar] [CrossRef]
  21. Kim, S.H.; Chung, J.M. An experimental model for peripheral neuropathy produced by segmental spinal nerve ligation in the rat. Pain 1992, 50, 355–363. [Google Scholar] [CrossRef]
  22. Bennett, G.J.; Xie, Y.K. A peripheral mononeuropathy in rat that produces disorders of pain sensation like those seen in man. Pain 1988, 33, 87–107. [Google Scholar] [CrossRef]
  23. Decosterd, I.; Woolf, C.J. Spared nerve injury: An animal model of persistent peripheral neuropathic pain. Pain 2000, 87, 149–158. [Google Scholar] [CrossRef]
  24. Challa, S.R. Surgical animal models of neuropathic pain: Pros and Cons. Int. J. Neurosci. 2015, 125, 170–174. [Google Scholar] [CrossRef] [PubMed]
  25. Tseng, T.J.; Chen, C.C.; Hsieh, Y.L.; Hsieh, S.T. Effects of decompression on neuropathic pain behaviors and skin reinnervation in chronic constriction injury. Exp. Neurol. 2007, 204, 574–582. [Google Scholar] [CrossRef]
  26. Tseng, T.J.; Hsieh, Y.L.; Hsieh, S.T. Reversal of ERK activation in the dorsal horn after decompression in chronic constriction injury. Exp. Neurol. 2007, 206, 17–23. [Google Scholar] [CrossRef] [PubMed]
  27. Chang, M.F.; Hsieh, J.H.; Chiang, H.; Kan, H.W.; Huang, C.M.; Chellis, L.; Lin, B.S.; Miaw, S.C.; Pan, C.L.; Chao, C.C.; et al. Effective gene expression in the rat dorsal root ganglia with a non-viral vector delivered via spinal nerve injection. Sci. Rep. 2016, 6, 35612. [Google Scholar] [CrossRef] [PubMed]
  28. Hovaguimian, A.; Gibbons, C.H. Diagnosis and treatment of pain in small-fiber neuropathy. Curr. Pain Headache Rep. 2011, 15, 193–200. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Cliffer, K.D.; Siuciak, J.A.; Carson, S.R.; Radley, H.E.; Park, J.S.; Lewis, D.R.; Zlotchenko, E.; Nguyen, T.; Garcia, K.; Tonra, J.R.; et al. Physiological characterization of Taxol-induced large-fiber sensory neuropathy in the rat. Ann. Neurol. 1998, 43, 46–55. [Google Scholar] [CrossRef]
  30. Davies, M.; Brophy, S.; Williams, R.; Taylor, A. The prevalence, severity, and impact of painful diabetic peripheral neuropathy in type 2 diabetes. Diabetes Care 2006, 29, 1518–1522. [Google Scholar] [CrossRef] [Green Version]
  31. Hsieh, Y.L.; Chiang, H.; Tseng, T.J.; Hsieh, S.T. Enhancement of cutaneous nerve regeneration by 4-methylcatechol in resiniferatoxin-induced neuropathy. J. Neuropathol. Exp. Neurol. 2008, 67, 93–104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Lee, Y.C.; Lu, S.C.; Hsieh, Y.L. Establishing a Mouse Model of a Pure Small Fiber Neuropathy with the Ultrapotent Agonist of Transient Receptor Potential Vanilloid Type 1. J. Vis. Exp. 2018. [Google Scholar] [CrossRef] [PubMed]
  33. Hsiao, T.H.; Fu, Y.S.; Ho, W.Y.; Chen, T.H.; Hsieh, Y.L. Promotion of thermal analgesia and neuropeptidergic skin reinnervation by 4-methylcatechol in resiniferatoxin-induced neuropathy. Kaohsiung J. Med. Sci. 2013, 29, 405–411. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Spencer, P.S.; Schaumburg, H.H. Nervous system degeneration produced by acrylamide monomer. Environ. Health Perspect. 1975, 11, 129–133. [Google Scholar] [CrossRef]
  35. Suzuki, K.; Pfaff, L.D. Acrylamide neuropathy in rats. An electron microscopic study of degeneration and regeneration. Acta Neuropathol. 1973, 24, 197–213. [Google Scholar] [CrossRef] [PubMed]
  36. Ko, M.H.; Chen, W.P.; Hsieh, S.T. Neuropathology of skin denervation in acrylamide-induced neuropathy. Neurobiol. Dis. 2002, 11, 155–165. [Google Scholar] [CrossRef] [Green Version]
  37. Ling, B.; Authier, N.; Balayssac, D.; Eschalier, A.; Coudore, F. Assessment of nociception in acrylamide-induced neuropathy in rats. Pain 2005, 119, 104–112. [Google Scholar] [CrossRef]
  38. Yousef, M.I.; El-Demerdash, F.M. Acrylamide-induced oxidative stress and biochemical perturbations in rats. Toxicology 2006, 219, 133–141. [Google Scholar] [CrossRef]
  39. Yousefsani, B.S.; Akbarizadeh, N.; Pourahmad, J. The antioxidant and neuroprotective effects of Zolpidem on acrylamide-induced neurotoxicity using Wistar rat primary neuronal cortical culture. Toxicol. Rep. 2020, 7, 233–240. [Google Scholar] [CrossRef]
  40. Oda, S.S. Metformin Protects against Experimental Acrylamide Neuropathy in Rats. Drug Dev. Res. 2017, 78, 349–359. [Google Scholar] [CrossRef]
  41. Seretny, M.; Currie, G.L.; Sena, E.S.; Ramnarine, S.; Grant, R.; MacLeod, M.R.; Colvin, L.A.; Fallon, M. Incidence, prevalence, and predictors of chemotherapy-induced peripheral neuropathy: A systematic review and meta-analysis. Pain 2014, 155, 2461–2470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Balayssac, D.; Ferrier, J.; Descoeur, J.; Ling, B.; Pezet, D.; Eschalier, A.; Authier, N. Chemotherapy-induced peripheral neuropathies: From clinical relevance to preclinical evidence. Expert Opin. Drug Saf. 2011, 10, 407–417. [Google Scholar] [CrossRef] [PubMed]
  43. Yang, I.H.; Siddique, R.; Hosmane, S.; Thakor, N.; Hoke, A. Compartmentalized microfluidic culture platform to study mechanism of paclitaxel-induced axonal degeneration. Exp. Neurol. 2009, 218, 124–128. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Ravula, S.K.; Wang, M.S.; McClain, M.A.; Asress, S.A.; Frazier, B.; Glass, J.D. Spatiotemporal localization of injury potentials in DRG neurons during vincristine-induced axonal degeneration. Neurosci. Lett. 2007, 415, 34–39. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. McDonald, E.S.; Windebank, A.J. Cisplatin-induced apoptosis of DRG neurons involves bax redistribution and cytochrome c release but not fas receptor signaling. Neurobiol. Dis. 2002, 9, 220–233. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Fukuda, Y.; Li, Y.; Segal, R.A. A Mechanistic Understanding of Axon Degeneration in Chemotherapy-Induced Peripheral Neuropathy. Front. Neurosci. 2017, 11, 481. [Google Scholar] [CrossRef] [PubMed]
  47. Boehmerle, W.; Huehnchen, P.; Peruzzaro, S.; Balkaya, M.; Endres, M. Electrophysiological, behavioral and histological characterization of paclitaxel, cisplatin, vincristine and bortezomib-induced neuropathy in C57Bl/6 mice. Sci. Rep. 2014, 4, 6370. [Google Scholar] [CrossRef] [PubMed]
  48. Chiang, H.Y.; Chien, H.F.; Shen, H.H.; Yang, J.D.; Chen, Y.H.; Chen, J.H.; Hsieh, S.T. Reinnervation of muscular targets by nerve regeneration through guidance conduits. J. Neuropathol. Exp. Neurol. 2005, 64, 576–587. [Google Scholar] [CrossRef] [Green Version]
  49. Chine, V.B.; Au, N.P.B.; Kumar, G.; Ma, C.H.E. Targeting Axon Integrity to Prevent Chemotherapy-Induced Peripheral Neuropathy. Mol. Neurobiol. 2019, 56, 3244–3259. [Google Scholar] [CrossRef]
  50. Calvo, M.; Davies, A.J.; Hebert, H.L.; Weir, G.A.; Chesler, E.J.; Finnerup, N.B.; Levitt, R.C.; Smith, B.H.; Neely, G.G.; Costigan, M.; et al. The Genetics of Neuropathic Pain from Model Organisms to Clinical Application. Neuron 2019, 104, 637–653. [Google Scholar] [CrossRef]
  51. Chen, L.; Liu, Y.W.; Yue, K.; Ru, Q.; Xiong, Q.; Ma, B.M.; Tian, X.; Li, C.Y. Differential expression of ATP-gated P2X receptors in DRG between chronic neuropathic pain and visceralgia rat models. Purinergic Signal. 2016, 12, 79–87. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Novakovic, S.D.; Kassotakis, L.C.; Oglesby, I.B.; Smith, J.A.; Eglen, R.M.; Ford, A.P.; Hunter, J.C. Immunocytochemical localization of P2X3 purinoceptors in sensory neurons in naive rats and following neuropathic injury. Pain 1999, 80, 273–282. [Google Scholar] [CrossRef]
  53. Leng, C.; Chen, L.; Li, C. Alteration of P2X1-6 receptor expression in retrograde Fluorogold-labeled DRG neurons from rat chronic neuropathic pain model. Biomed. Rep. 2019, 10, 225–230. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Barclay, J.; Patel, S.; Dorn, G.; Wotherspoon, G.; Moffatt, S.; Eunson, L.; Abdel’al, S.; Natt, F.; Hall, J.; Winter, J.; et al. Functional downregulation of P2X3 receptor subunit in rat sensory neurons reveals a significant role in chronic neuropathic and inflammatory pain. J. Neurosci. 2002, 22, 8139–8147. [Google Scholar] [CrossRef] [Green Version]
  55. Kage, K.; Niforatos, W.; Zhu, C.Z.; Lynch, K.J.; Honore, P.; Jarvis, M.F. Alteration of dorsal root ganglion P2X3 receptor expression and function following spinal nerve ligation in the rat. Exp. Brain Res. 2002, 147, 511–519. [Google Scholar] [CrossRef]
  56. Fukuoka, T.; Tokunaga, A.; Tachibana, T.; Dai, Y.; Yamanaka, H.; Noguchi, K. VR1, but not P2X(3), increases in the spared L4 DRG in rats with L5 spinal nerve ligation. Pain 2002, 99, 111–120. [Google Scholar] [CrossRef]
  57. Kim, C.; Chung, J.M.; Chung, K. Changes in the gene expression of six subtypes of P2X receptors in rat dorsal root ganglion after spinal nerve ligation. Neurosci. Lett. 2003, 337, 81–84. [Google Scholar] [CrossRef]
  58. Wang, Y.; Zhang, X.; Guo, Q.L.; Zou, W.Y.; Huang, C.S.; Yan, J.Q. Cyclooxygenase inhibitors suppress the expression of P2X(3) receptors in the DRG and attenuate hyperalgesia following chronic constriction injury in rats. Neurosci. Lett. 2010, 478, 77–81. [Google Scholar] [CrossRef]
  59. Hsieh, Y.L.; Chiang, H.; Lue, J.H.; Hsieh, S.T. P2X3-mediated peripheral sensitization of neuropathic pain in resiniferatoxin-induced neuropathy. Exp. Neurol. 2012, 235, 316–325. [Google Scholar] [CrossRef]
  60. Hori, K.; Ozaki, N.; Suzuki, S.; Sugiura, Y. Upregulations of P2X(3) and ASIC3 involve in hyperalgesia induced by cisplatin administration in rats. Pain 2010, 149, 393–405. [Google Scholar] [CrossRef]
  61. Watanabe, M.; Ueda, T.; Shibata, Y.; Kumamoto, N.; Ugawa, S. The role of TRPV1 channels in carrageenan-induced mechanical hyperalgesia in mice. Neuroreport 2015, 26, 173–178. [Google Scholar] [CrossRef] [PubMed]
  62. Caterina, M.J.; Leffler, A.; Malmberg, A.B.; Martin, W.J.; Trafton, J.; Petersen-Zeitz, K.R.; Koltzenburg, M.; Basbaum, A.I.; Julius, D. Impaired nociception and pain sensation in mice lacking the capsaicin receptor. Science 2000, 288, 306–313. [Google Scholar] [CrossRef] [PubMed]
  63. Story, G.M.; Peier, A.M.; Reeve, A.J.; Eid, S.R.; Mosbacher, J.; Hricik, T.R.; Earley, T.J.; Hergarden, A.C.; Andersson, D.A.; Hwang, S.W.; et al. ANKTM1, a TRP-like channel expressed in nociceptive neurons, is activated by cold temperatures. Cell 2003, 112, 819–829. [Google Scholar] [CrossRef] [Green Version]
  64. Su, L.; Shu, R.; Song, C.; Yu, Y.; Wang, G.; Li, Y.; Liu, C. Downregulations of TRPM8 expression and membrane trafficking in dorsal root ganglion mediate the attenuation of cold hyperalgesia in CCI rats induced by GFRalpha3 knockdown. Brain Res. Bull. 2017, 135, 8–24. [Google Scholar] [CrossRef]
  65. Dib-Hajj, S.D.; Cummins, T.R.; Black, J.A.; Waxman, S.G. Sodium channels in normal and pathological pain. Annu. Rev. Neurosci. 2010, 33, 325–347. [Google Scholar] [CrossRef] [Green Version]
  66. Su, L.; Wang, C.; Yu, Y.H.; Ren, Y.Y.; Xie, K.L.; Wang, G.L. Role of TRPM8 in dorsal root ganglion in nerve injury-induced chronic pain. BMC Neurosci. 2011, 12, 120. [Google Scholar] [CrossRef] [Green Version]
  67. Guo, S.H.; Lin, J.P.; Huang, L.E.; Yang, Y.; Chen, C.Q.; Li, N.N.; Su, M.Y.; Zhao, X.; Zhu, S.M.; Yao, Y.X. Silencing of spinal Trpv1 attenuates neuropathic pain in rats by inhibiting CAMKII expression and ERK2 phosphorylation. Sci. Rep. 2019, 9, 2769. [Google Scholar] [CrossRef] [Green Version]
  68. Black, J.A.; Dib-Hajj, S.; McNabola, K.; Jeste, S.; Rizzo, M.A.; Kocsis, J.D.; Waxman, S.G. Spinal sensory neurons express multiple sodium channel alpha-subunit mRNAs. Brain Res. Mol. Brain Res. 1996, 43, 117–131. [Google Scholar] [CrossRef]
  69. Dib-Hajj, S.D.; Tyrrell, L.; Black, J.A.; Waxman, S.G. NaN, a novel voltage-gated Na channel, is expressed preferentially in peripheral sensory neurons and down-regulated after axotomy. Proc. Natl. Acad. Sci. USA 1998, 95, 8963–8968. [Google Scholar] [CrossRef] [Green Version]
  70. Waxman, S.G. Transcriptional channelopathies: An emerging class of disorders. Nat. Rev. Neurosci. 2001, 2, 652–659. [Google Scholar] [CrossRef]
  71. Chattopadhyay, M.; Mata, M.; Fink, D.J. Continuous delta-opioid receptor activation reduces neuronal voltage-gated sodium channel (NaV1.7) levels through activation of protein kinase C in painful diabetic neuropathy. J. Neurosci. 2008, 28, 6652–6658. [Google Scholar] [CrossRef] [PubMed]
  72. Hong, S.; Morrow, T.J.; Paulson, P.E.; Isom, L.L.; Wiley, J.W. 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. 2004, 279, 29341–29350. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Laedermann, C.J.; Cachemaille, M.; Kirschmann, G.; Pertin, M.; Gosselin, R.D.; Chang, I.; Albesa, M.; Towne, C.; Schneider, B.L.; Kellenberger, S.; et al. Dysregulation of voltage-gated sodium channels by ubiquitin ligase NEDD4-2 in neuropathic pain. J. Clin. Investig. 2013, 123, 3002–3013. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Catterall, W.A.; Perez-Reyes, E.; Snutch, T.P.; Striessnig, J. International Union of Pharmacology. XLVIII. Nomenclature and structure-function relationships of voltage-gated calcium channels. Pharmacol. Rev. 2005, 57, 411–425. [Google Scholar] [CrossRef] [PubMed]
  75. Iftinca, M.C. Neuronal T-type calcium channels: What’s new? Iftinca: T-type channel regulation. J. Med. Life 2011, 4, 126–138. [Google Scholar] [PubMed]
  76. Dolphin, A.C. Voltage-gated calcium channel alpha 2delta subunits: An assessment of proposed novel roles. F1000Res 2018, 7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Talley, E.M.; Cribbs, L.L.; Lee, J.H.; Daud, A.; Perez-Reyes, E.; Bayliss, D.A. Differential distribution of three members of a gene family encoding low voltage-activated (T-type) calcium channels. J. Neurosci. 1999, 19, 1895–1911. [Google Scholar] [CrossRef] [PubMed]
  78. Shin, J.B.; Martinez-Salgado, C.; Heppenstall, P.A.; Lewin, G.R. A T-type calcium channel required for normal function of a mammalian mechanoreceptor. Nat. Neurosci. 2003, 6, 724–730. [Google Scholar] [CrossRef]
  79. Rose, K.E.; Lunardi, N.; Boscolo, A.; Dong, X.; Erisir, A.; Jevtovic-Todorovic, V.; Todorovic, S.M. Immunohistological demonstration of CaV3.2 T-type voltage-gated calcium channel expression in soma of dorsal root ganglion neurons and peripheral axons of rat and mouse. Neuroscience 2013, 250, 263–274. [Google Scholar] [CrossRef] [Green Version]
  80. Kang, X.J.; Chi, Y.N.; Chen, W.; Liu, F.Y.; Cui, S.; Liao, F.F.; Cai, J.; Wan, Y. Increased expression of CaV3.2 T-type calcium channels in damaged DRG neurons contributes to neuropathic pain in rats with spared nerve injury. Mol. Pain 2018, 14. [Google Scholar] [CrossRef] [Green Version]
  81. Luo, Z.D.; Chaplan, S.R.; Higuera, E.S.; Sorkin, L.S.; Stauderman, K.A.; Williams, M.E.; Yaksh, T.L. Upregulation of dorsal root ganglion (alpha)2(delta) calcium channel subunit and its correlation with allodynia in spinal nerve-injured rats. J. Neurosci. 2001, 21, 1868–1875. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Kim, D.S.; Yoon, C.H.; Lee, S.J.; Park, S.Y.; Yoo, H.J.; Cho, H.J. Changes in voltage-gated calcium channel alpha(1) gene expression in rat dorsal root ganglia following peripheral nerve injury. Brain Res. Mol. Brain Res. 2001, 96, 151–156. [Google Scholar] [CrossRef]
  83. Lotem, J.; Sachs, L. Epigenetics and the plasticity of differentiation in normal and cancer stem cells. Oncogene 2006, 25, 7663–7672. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Atlasi, Y.; Stunnenberg, H.G. The interplay of epigenetic marks during stem cell differentiation and development. Nat. Rev. Genet. 2017, 18, 643–658. [Google Scholar] [CrossRef]
  85. Penas, C.; Navarro, X. Epigenetic Modifications Associated to Neuroinflammation and Neuropathic Pain After Neural Trauma. Front. Cell. Neurosci. 2018, 12, 158. [Google Scholar] [CrossRef] [Green Version]
  86. Cheng, H.Y.; Penninger, J.M. Transcriptional mechanisms underlying neuropathic pain: DREAM, transcription factors and future pain management? Expert Rev. Neurother. 2002, 2, 677–689. [Google Scholar] [CrossRef]
  87. Moss, A.; Blackburn-Munro, G.; Garry, E.M.; Blakemore, J.A.; Dickinson, T.; Rosie, R.; Mitchell, R.; Fleetwood-Walker, S.M. A role of the ubiquitin-proteasome system in neuropathic pain. J. Neurosci. 2002, 22, 1363–1372. [Google Scholar] [CrossRef] [Green Version]
  88. Vogenberg, F.R.; Isaacson Barash, C.; Pursel, M. Personalized medicine: Part 1: Evolution and development into theranostics. Pharm. Ther. 2010, 35, 560–576. [Google Scholar]
  89. Khorasanizadeh, S. The nucleosome: From genomic organization to genomic regulation. Cell 2004, 116, 259–272. [Google Scholar] [CrossRef] [Green Version]
  90. Jenuwein, T.; Allis, C.D. Translating the histone code. Science 2001, 293, 1074–1080. [Google Scholar] [CrossRef] [Green Version]
  91. Shao, C.; Gao, Y.; Jin, D.; Xu, X.; Tan, S.; Yu, H.; Zhao, Q.; Zhao, L.; Wang, W.; Wang, D. DNMT3a methylation in neuropathic pain. J. Pain Res. 2017, 10, 2253–2262. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Garriga, J.; Laumet, G.; Chen, S.R.; Zhang, Y.; Madzo, J.; Issa, J.J.; Pan, H.L.; Jelinek, J. Nerve Injury-Induced Chronic Pain Is Associated with Persistent DNA Methylation Reprogramming in Dorsal Root Ganglion. J. Neurosci. 2018, 38, 6090–6101. [Google Scholar] [CrossRef] [PubMed]
  93. Zhang, H.H.; Hu, J.; Zhou, Y.L.; Qin, X.; Song, Z.Y.; Yang, P.P.; Hu, S.; Jiang, X.; Xu, G.Y. Promoted Interaction of Nuclear Factor-kappaB With Demethylated Purinergic P2X3 Receptor Gene Contributes to Neuropathic Pain in Rats With Diabetes. Diabetes 2015, 64, 4272–4284. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Zhao, J.Y.; Liang, L.; Gu, X.; Li, Z.; Wu, S.; Sun, L.; Atianjoh, F.E.; Feng, J.; Mo, K.; Jia, S.; et al. DNA methyltransferase DNMT3a contributes to neuropathic pain by repressing Kcna2 in primary afferent neurons. Nat. Commun. 2017, 8, 14712. [Google Scholar] [CrossRef] [PubMed]
  95. Chuang, D.M.; Leng, Y.; Marinova, Z.; Kim, H.J.; Chiu, C.T. Multiple roles of HDAC inhibition in neurodegenerative conditions. Trends Neurosci. 2009, 32, 591–601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Guan, J.S.; Haggarty, S.J.; Giacometti, E.; Dannenberg, J.H.; Joseph, N.; Gao, J.; Nieland, T.J.; Zhou, Y.; Wang, X.; Mazitschek, R.; et al. HDAC2 negatively regulates memory formation and synaptic plasticity. Nature 2009, 459, 55–60. [Google Scholar] [CrossRef]
  97. Sando, R., III; Gounko, N.; Pieraut, S.; Liao, L.; Yates, J., III; Maximov, A. HDAC4 governs a transcriptional program essential for synaptic plasticity and memory. Cell 2012, 151, 821–834. [Google Scholar] [CrossRef] [Green Version]
  98. Gu, P.; Pan, Z.; Wang, X.M.; Sun, L.; Tai, L.W.; Cheung, C.W. Histone deacetylase 5 (HDAC5) regulates neuropathic pain through SRY-related HMG-box 10 (SOX10)-dependent mechanism in mice. Pain 2018, 159, 526–539. [Google Scholar] [CrossRef]
  99. Matsushita, Y.; Araki, K.; Omotuyi, O.; Mukae, T.; Ueda, H. HDAC inhibitors restore C-fibre sensitivity in experimental neuropathic pain model. Br. J. Pharmacol. 2013, 170, 991–998. [Google Scholar] [CrossRef] [Green Version]
  100. Denk, F.; Huang, W.; Sidders, B.; Bithell, A.; Crow, M.; Grist, J.; Sharma, S.; Ziemek, D.; Rice, A.S.; Buckley, N.J.; et al. HDAC inhibitors attenuate the development of hypersensitivity in models of neuropathic pain. Pain 2013, 154, 1668–1679. [Google Scholar] [CrossRef] [Green Version]
  101. Ji, R.R.; Baba, H.; Brenner, G.J.; Woolf, C.J. Nociceptive-specific activation of ERK in spinal neurons contributes to pain hypersensitivity. Nat. Neurosci. 1999, 2, 1114–1119. [Google Scholar] [CrossRef] [PubMed]
  102. Samour, M.S.; Shaikh, S.M.; Mahns, D.A.; Shortland, P.J. Noxious, but not innocuous, thermal stimuli evoke pERK expression in dorsal horn neurons after spared nerve injury in adult rats. Neurosci. Lett. 2017, 654, 49–55. [Google Scholar] [CrossRef] [PubMed]
  103. Qu, Y.J.; Jia, L.; Zhang, X.; Wei, H.; Yue, S.W. MAPK Pathways Are Involved in Neuropathic Pain in Rats with Chronic Compression of the Dorsal Root Ganglion. Evid. Based Complement. Alternat. Med. 2016, 2016, 6153215. [Google Scholar] [CrossRef] [PubMed]
  104. Chuderland, D.; Konson, A.; Seger, R. Identification and characterization of a general nuclear translocation signal in signaling proteins. Mol. Cell 2008, 31, 850–861. [Google Scholar] [CrossRef] [PubMed]
  105. Impey, S.; Obrietan, K.; Wong, S.T.; Poser, S.; Yano, S.; Wayman, G.; Deloulme, J.C.; Chan, G.; Storm, D.R. Cross talk between ERK and PKA is required for Ca2+ stimulation of CREB-dependent transcription and ERK nuclear translocation. Neuron 1998, 21, 869–883. [Google Scholar] [CrossRef] [Green Version]
  106. Monje, P.; Hernandez-Losa, J.; Lyons, R.J.; Castellone, M.D.; Gutkind, J.S. Regulation of the transcriptional activity of c-Fos by ERK. A novel role for the prolyl isomerase PIN1. J. Biol. Chem. 2005, 280, 35081–35084. [Google Scholar] [CrossRef] [Green Version]
  107. Gao, Y.J.; Ji, R.R. c-Fos and pERK, which is a better marker for neuronal activation and central sensitization after noxious stimulation and tissue injury? Open Pain J. 2009, 2, 11–17. [Google Scholar] [CrossRef]
  108. Seijffers, R.; Mills, C.D.; Woolf, C.J. ATF3 increases the intrinsic growth state of DRG neurons to enhance peripheral nerve regeneration. J. Neurosci. 2007, 27, 7911–7920. [Google Scholar] [CrossRef] [Green Version]
  109. Tsujino, H.; Kondo, E.; Fukuoka, T.; Dai, Y.; Tokunaga, A.; Miki, K.; Yonenobu, K.; Ochi, T.; Noguchi, K. Activating transcription factor 3 (ATF3) induction by axotomy in sensory and motoneurons: A novel neuronal marker of nerve injury. Mol. Cell. Neurosci. 2000, 15, 170–182. [Google Scholar] [CrossRef]
  110. Ding, S.; Yu, Q.; Wang, J.; Zhu, L.; Li, T.; Guo, X.; Zhang, X. Activation of ATF3/AP-1 signaling pathway is required for P2X3-induced endometriosis pain. Hum. Reprod. 2020, 35, 1130–1144. [Google Scholar] [CrossRef]
  111. Staub, O.; Dho, S.; Henry, P.; Correa, J.; Ishikawa, T.; McGlade, J.; Rotin, D. WW domains of Nedd4 bind to the proline-rich PY motifs in the epithelial Na+ channel deleted in Liddle’s syndrome. EMBO J. 1996, 15, 2371–2380. [Google Scholar] [CrossRef] [PubMed]
  112. Younger, J.M.; Ren, H.Y.; Chen, L.; Fan, C.Y.; Fields, A.; Patterson, C.; Cyr, D.M. A foldable CFTR{Delta}F508 biogenic intermediate accumulates upon inhibition of the Hsc70-CHIP E3 ubiquitin ligase. J. Cell Biol. 2004, 167, 1075–1085. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Altier, C.; Garcia-Caballero, A.; Simms, B.; You, H.; Chen, L.; Walcher, J.; Tedford, H.W.; Hermosilla, T.; Zamponi, G.W. The Cavbeta subunit prevents RFP2-mediated ubiquitination and proteasomal degradation of L-type channels. Nat. Neurosci. 2011, 14, 173–180. [Google Scholar] [CrossRef] [PubMed]
  114. Garcia-Caballero, A.; Gadotti, V.M.; Stemkowski, P.; Weiss, N.; Souza, I.A.; Hodgkinson, V.; Bladen, C.; Chen, L.; Hamid, J.; Pizzoccaro, A.; et al. The deubiquitinating enzyme USP5 modulates neuropathic and inflammatory pain by enhancing Cav3.2 channel activity. Neuron 2014, 83, 1144–1158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Nascimento, A.I.; Mar, F.M.; Sousa, M.M. The intriguing nature of dorsal root ganglion neurons: Linking structure with polarity and function. Prog. Neurobiol. 2018, 168, 86–103. [Google Scholar] [CrossRef] [PubMed]
  116. Haberberger, R.V.; Barry, C.; Dominguez, N.; Matusica, D. Human Dorsal Root Ganglia. Front. Cell. Neurosci. 2019, 13, 271. [Google Scholar] [CrossRef] [PubMed]
  117. Puljak, L.; Kojundzic, S.L.; Hogan, Q.H.; Sapunar, D. Lidocaine injection into the rat dorsal root ganglion causes neuroinflammation. Anesth. Analg. 2009, 108, 1021–1026. [Google Scholar] [CrossRef] [Green Version]
  118. Fischer, G.; Kostic, S.; Nakai, H.; Park, F.; Sapunar, D.; Yu, H.; Hogan, Q. Direct injection into the dorsal root ganglion: Technical, behavioral, and histological observations. J. Neurosci. Methods 2011, 199, 43–55. [Google Scholar] [CrossRef] [Green Version]
  119. Xu, J.J.; Walla, B.C.; Diaz, M.F.; Fuller, G.N.; Gutstein, H.B. Intermittent lumbar puncture in rats: A novel method for the experimental study of opioid tolerance. Anesth. Analg. 2006, 103, 714–720. [Google Scholar] [CrossRef]
  120. Haggerty, D.L.; Grecco, G.G.; Reeves, K.C.; Atwood, B. Adeno-Associated Viral Vectors in Neuroscience Research. Mol. Ther. Methods Clin. Dev. 2020, 17, 69–82. [Google Scholar] [CrossRef]
  121. Daya, S.; Berns, K.I. Gene therapy using adeno-associated virus vectors. Clin. Microbiol. Rev. 2008, 21, 583–593. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Mason, M.R.; Ehlert, E.M.; Eggers, R.; Pool, C.W.; Hermening, S.; Huseinovic, A.; Timmermans, E.; Blits, B.; Verhaagen, J. Comparison of AAV serotypes for gene delivery to dorsal root ganglion neurons. Mol. Ther. 2010, 18, 715–724. [Google Scholar] [CrossRef] [PubMed]
  123. Haery, L.; Deverman, B.E.; Matho, K.S.; Cetin, A.; Woodard, K.; Cepko, C.; Guerin, K.I.; Rego, M.A.; Ersing, I.; Bachle, S.M.; et al. Adeno-Associated Virus Technologies and Methods for Targeted Neuronal Manipulation. Front. Neuroanat. 2019, 13, 93. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Mingozzi, F.; High, K.A. Immune responses to AAV vectors: Overcoming barriers to successful gene therapy. Blood 2013, 122, 23–36. [Google Scholar] [CrossRef]
  125. Russell, S.; Bennett, J.; Wellman, J.A.; Chung, D.C.; Yu, Z.F.; Tillman, A.; Wittes, J.; Pappas, J.; Elci, O.; McCague, S.; et al. Efficacy and safety of voretigene neparvovec (AAV2-hRPE65v2) in patients with RPE65-mediated inherited retinal dystrophy: A randomised, controlled, open-label, phase 3 trial. Lancet 2017, 390, 849–860. [Google Scholar] [CrossRef]
  126. Kushnarev, M.; Pirvulescu, I.P.; Candido, K.D.; Knezevic, N.N. Neuropathic pain: Preclinical and early clinical progress with voltage-gated sodium channel blockers. Expert Opin. Investig. Drugs 2020, 29, 259–271. [Google Scholar] [CrossRef]
  127. Urru, M.; Muzzi, M.; Coppi, E.; Ranieri, G.; Buonvicino, D.; Camaioni, E.; Coppini, R.; Pugliese, A.M.; Tanaka, B.; Estacion, M.; et al. Dexpramipexole blocks Nav1.8 sodium channels and provides analgesia in multiple nociceptive and neuropathic pain models. Pain 2020, 161, 831–841. [Google Scholar] [CrossRef]
  128. Cudkowicz, M.; Bozik, M.E.; Ingersoll, E.W.; Miller, R.; Mitsumoto, H.; Shefner, J.; Moore, D.H.; Schoenfeld, D.; Mather, J.L.; Archibald, D.; et al. The effects of dexpramipexole (KNS-760704) in individuals with amyotrophic lateral sclerosis. Nat. Med. 2011, 17, 1652–1656. [Google Scholar] [CrossRef]
  129. Cudkowicz, M.E.; van den Berg, L.H.; Shefner, J.M.; Mitsumoto, H.; Mora, J.S.; Ludolph, A.; Hardiman, O.; Bozik, M.E.; Ingersoll, E.W.; Archibald, D.; et al. Dexpramipexole versus placebo for patients with amyotrophic lateral sclerosis (EMPOWER): A randomised, double-blind, phase 3 trial. Lancet Neurol. 2013, 12, 1059–1067. [Google Scholar] [CrossRef]
  130. Dworetzky, S.I.; Hebrank, G.T.; Archibald, D.G.; Reynolds, I.J.; Farwell, W.; Bozik, M.E. The targeted eosinophil-lowering effects of dexpramipexole in clinical studies. Blood Cells Mol. Dis. 2017, 63, 62–65. [Google Scholar] [CrossRef]
  131. Ji, G.; Zhou, S.; Carlton, S.M. Intact Adelta-fibers up-regulate transient receptor potential A1 and contribute to cold hypersensitivity in neuropathic rats. Neuroscience 2008, 154, 1054–1066. [Google Scholar] [CrossRef] [Green Version]
  132. Li, M.; Zhang, S.J.; Yang, L.; Fang, X.L.; Hu, H.F.; Zhao, M.Y.; Li, L.; Guo, Y.Y.; Shao, J.P. Voltage-gated sodium channel 1.7 expression decreases in dorsal root ganglia in a spinal nerve ligation neuropathic pain model. Kaohsiung J. Med. Sci. 2019, 35, 493–500. [Google Scholar] [CrossRef]
  133. Ma, R.S.Y.; Kayani, K.; Whyte-Oshodi, D.; Whyte-Oshodi, A.; Nachiappan, N.; Gnanarajah, S.; Mohammed, R. Voltage gated sodium channels as therapeutic targets for chronic pain. J. Pain Res. 2019, 12, 2709–2722. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Chen, L.; Leng, C.; Ru, Q.; Xiong, Q.; Zhou, M.; Wu, Y. Retrograde Labeling of Different Distribution Features of DRG P2X2 and P2X3 Receptors in a Neuropathic Pain Rat Model. Biomed. Res. Int. 2020, 2020, 9861459. [Google Scholar] [CrossRef] [PubMed]
  135. Chen, Y.; Li, G.W.; Wang, C.; Gu, Y.; Huang, L.Y. Mechanisms underlying enhanced P2X receptor-mediated responses in the neuropathic pain state. Pain 2005, 119, 38–48. [Google Scholar] [CrossRef] [PubMed]
  136. Frederick, J.; Buck, M.E.; Matson, D.J.; Cortright, D.N. Increased TRPA1, TRPM8, and TRPV2 expression in dorsal root ganglia by nerve injury. Biochem. Biophys. Res. Commun. 2007, 358, 1058–1064. [Google Scholar] [CrossRef] [PubMed]
  137. Wang, Z.; Ling, D.; Wu, C.; Han, J.; Zhao, Y. Baicalin prevents the up-regulation of TRPV1 in dorsal root ganglion and attenuates chronic neuropathic pain. Vet. Med. Sci. 2020. [Google Scholar] [CrossRef] [PubMed]
  138. Staaf, S.; Oerther, S.; Lucas, G.; Mattsson, J.P.; Ernfors, P. Differential regulation of TRP channels in a rat model of neuropathic pain. Pain 2009, 144, 187–199. [Google Scholar] [CrossRef] [PubMed]
  139. Casals-Diaz, L.; Casas, C.; Navarro, X. Changes of voltage-gated sodium channels in sensory nerve regeneration and neuropathic pain models. Restor. Neurol. Neurosci. 2015, 33, 321–334. [Google Scholar] [CrossRef]
  140. Laedermann, C.J.; Pertin, M.; Suter, M.R.; Decosterd, I. Voltage-gated sodium channel expression in mouse DRG after SNI leads to re-evaluation of projections of injured fibers. Mol. Pain 2014, 10, 19. [Google Scholar] [CrossRef] [Green Version]
  141. Kan, H.W.; Chang, C.H.; Lin, C.L.; Lee, Y.C.; Hsieh, S.T.; Hsieh, Y.L. Downregulation of adenosine and adenosine A1 receptor contributes to neuropathic pain in resiniferatoxin neuropathy. Pain 2018, 159, 1580–1591. [Google Scholar] [CrossRef] [PubMed]
  142. Ta, L.E.; Bieber, A.J.; Carlton, S.M.; Loprinzi, C.L.; Low, P.A.; Windebank, A.J. Transient Receptor Potential Vanilloid 1 is essential for cisplatin-induced heat hyperalgesia in mice. Mol. Pain 2010, 6, 15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Leo, M.; Schmitt, L.I.; Jastrow, H.; Thomale, J.; Kleinschnitz, C.; Hagenacker, T. Cisplatin alters the function and expression of N-type voltage-gated calcium channels in the absence of morphological damage of sensory neurons. Mol. Pain 2017, 13. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic diagram illustrating the peripheral sensitization in resiniferatoxin (RTX)-induced neuropathic pain model. Thermal hypoalgesia and mechanical allodynia were observed after 1 week of RTX treatment. After RTX exposure, several dorsal root ganglia (DRG) profiles were changed including the increase of purinergic receptor P2X3 (P2X3)(+)/activating transcription factor 3 (ATF3)(+) neurons and the decrease of transient receptor potential vanilloid subtype 1 (TRPV1)(+)/calcitonin gene-related peptide (CGRP)(+) neurons. At the terminal ending of nerve, the phenotype of dermal nerve also had similar changes as the DRG. The number of CGRP(+) and P2X3(+) nerve were decreased and increased respectively.
Figure 1. Schematic diagram illustrating the peripheral sensitization in resiniferatoxin (RTX)-induced neuropathic pain model. Thermal hypoalgesia and mechanical allodynia were observed after 1 week of RTX treatment. After RTX exposure, several dorsal root ganglia (DRG) profiles were changed including the increase of purinergic receptor P2X3 (P2X3)(+)/activating transcription factor 3 (ATF3)(+) neurons and the decrease of transient receptor potential vanilloid subtype 1 (TRPV1)(+)/calcitonin gene-related peptide (CGRP)(+) neurons. At the terminal ending of nerve, the phenotype of dermal nerve also had similar changes as the DRG. The number of CGRP(+) and P2X3(+) nerve were decreased and increased respectively.
Cells 09 02725 g001
Table 1. Injury-induced neuropathic pain models.
Table 1. Injury-induced neuropathic pain models.
AdvantageDisadvantageApplication
SNL [21]Specific injury in L5 and L6 DRGTime consumingSpinal nerve injection
CCI [22]Combine injury and uninjured fibersVariation in each operationDecompression model
SNI [23]Combine injury and uninjured fibers
Robustness of the response
Low local inflammation
Abbreviation: SNL, spinal nerve ligation; CCI, chronic constriction injury; and SNI, spared nerve injury.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations

Share and Cite

MDPI and ACS Style

Yeh, T.-Y.; Luo, I.-W.; Hsieh, Y.-L.; Tseng, T.-J.; Chiang, H.; Hsieh, S.-T. Peripheral Neuropathic Pain: From Experimental Models to Potential Therapeutic Targets in Dorsal Root Ganglion Neurons. Cells 2020, 9, 2725. https://doi.org/10.3390/cells9122725

AMA Style

Yeh T-Y, Luo I-W, Hsieh Y-L, Tseng T-J, Chiang H, Hsieh S-T. Peripheral Neuropathic Pain: From Experimental Models to Potential Therapeutic Targets in Dorsal Root Ganglion Neurons. Cells. 2020; 9(12):2725. https://doi.org/10.3390/cells9122725

Chicago/Turabian Style

Yeh, Ti-Yen, I-Wei Luo, Yu-Lin Hsieh, To-Jung Tseng, Hao Chiang, and Sung-Tsang Hsieh. 2020. "Peripheral Neuropathic Pain: From Experimental Models to Potential Therapeutic Targets in Dorsal Root Ganglion Neurons" Cells 9, no. 12: 2725. https://doi.org/10.3390/cells9122725

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop