Next Article in Journal
Early-Life Sublethal Exposure to Thiacloprid Alters Adult Honeybee Gut Microbiota
Previous Article in Journal
Genome-Wide Identification and Characterization of Ammonium Transporter (AMT) Genes in Chlamydomonas reinhardtii
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Therapeutic Implication of miRNAs as an Active Regulatory Player in the Management of Pain: A Review

by
Mubashir Hassan
1,*,
Saba Shahzadi
1,
Muhammad Yasir
2,
Wanjoo Chun
2 and
Andrzej Kloczkowski
1,3,4
1
The Steve and Cindy Rasmussen Institute for Genomic Medicine at Nationwide Children’s Hospital, Columbus, OH 43205, USA
2
Department of Pharmacology, College of Medicine, Kangwon National University, Chuncheon 24341, Republic of Korea
3
Department of Pediatrics, The Ohio State University School of Medicine, Columbus, OH 43205, USA
4
Department of Biomedical Informatics, The Ohio State University, Columbus, OH 43210, USA
*
Author to whom correspondence should be addressed.
Genes 2024, 15(8), 1003; https://doi.org/10.3390/genes15081003
Submission received: 30 November 2023 / Revised: 15 July 2024 / Accepted: 26 July 2024 / Published: 31 July 2024
(This article belongs to the Section RNA)

Abstract

:
Chronic pain is frequently associated with neuropathy, inflammation, or the malfunctioning of nerves. Chronic pain is associated with a significant burden of morbidity due to opioid use, associated with addiction and tolerance, and disability. MicroRNAs (miRs) are emerging therapeutic targets to treat chronic pain through the regulation of genes associated with inflammation, neuronal excitability, survival, or de-differentiation. In this review, we discuss the possible involvement of miRs in pain-related molecular pathways. miRs are known to regulate high-conviction pain genes, supporting their potential as therapeutic targets.

1. Introduction

Pain receptors, also known as nociceptors, are specialized nerves that detect and transmit signals in response to potentially harmful stimuli such as temperature extremes, tissue injury, pressure, or irritants [1,2,3]. Upon the activation of pain receptors by a noxious stimulus, they transmit action potentials to spinal interneurons and, in turn, to the brain, where the signal may be interpreted as pain [4,5]. This initiates aversive responses such as pulling away from a hot surface or conscious perceptions of pain [6]. Different subtypes of nociceptors respond preferentially to mechanical, thermal, or chemical stimuli [7]. Furthermore, activation of nociceptors triggers inflammatory responses that both sensitize nociceptors and contribute to the perception of pain.
Generally, nociceptors express membrane receptors that detect heat, cold, or different types of hazardous substances [8]. Activation of these receptors promotes nociceptive action potentials, characteristically in unmyelinated C-fibers and myelinated Aδ-fibers [9,10], and releases excitatory neurotransmitters onto interneurons in the spine (especially the dorsal horn) [9,10]. The spinal dorsal horn’s excitatory and inhibitory interneurons can influence the perception of pain associated with nociceptive active potentials [11,12]. These physiological pain responses protect against tissue damage by triggering aversive behaviors. However, increased activity and responses of spinal dorsal horn neurons—a phenomenon known as central sensitization—are frequently linked to persistent pathological pain, including neuropathic and inflammatory pain [13,14]. Central sensitization forms the cellular basis for hyperalgesia (extreme sensitivity to pain) and allodynia (pain due to a stimulus that does not normally provoke pain) [15,16]. Alternatively, chronic peripheral activation of nociceptors can also lead to chronic pain, a condition that can be difficult to treat and can have a significant impact on quality of life.
MicroRNAs (miRs) are non-coding, tiny RNA molecules that control post-transcriptional regulation of gene expression. The biosynthesis of miRs is divided into two main steps: transcription and processing [17,18]. The first step in miR biosynthesis is the transcription of the DNA sequence encoding the miR by RNA polymerase II to produce a primary miR (pri-miR) transcript. Pri-miRNAs are typically several hundred nucleotides long and contain a stem-loop structure, with one end of the stem-loop being a single-stranded RNA tail [19]. The pri-miR transcript is processed by two RNase III enzymes, Drosha and Dicer, to generate the mature miRNA. First, the pri-miRNA is cleaved by Drosha in the nucleus to produce a precursor miR (pre-miR), which is typically 70–100 nucleotides long and contains a hairpin structure. After being transferred to the cytoplasm, the pre-miR undergoes further processing by Dicer to create a double-stranded RNA duplex. This duplex’s other strand, referred to as the passenger strand, is often eliminated while the other strand becomes the mature miR and is integrated into the RISC [20].
Pain management often utilizes medication to alleviate or control pain [21]. Nonsteroidal anti-inflammatory drugs (NSAIDs), such as aspirin, ibuprofen, and naproxen, are usually used as first-line therapy [22,23]. NSAIDs are associated with gastrointestinal issues, stomach irritation, ulcers, and gastrointestinal bleeding [24]. Furthermore, prolonged use of NSAIDs may increase the risk of heart attacks and strokes [25]. Finally, NSAIDs can also lead to kidney problems, particularly in individuals with pre-existing kidney disease [26]. These adverse effects both contribute to population-wide morbidity due to the wide use of NSAIDs and limit the use of NSAIDs in multimorbid patients [27].
Opioids, including oxycodone, hydrocodone, and morphine, are also sometimes used [28]. However, these drugs have a high potential for abuse, addiction, and physical dependence. Moreover, high doses or misuse can lead to slowed breathing and respiratory depression. Opioids also cause constipation and may cause drowsiness and impair cognitive function [29,30]. Finally, while not the sole cause, prescription opioids do contribute to the opioid overdose crisis [31]. Acetaminophen is an effective mild analgesic and antipyretic agent frequently used for pain management [32]. Reports shows that acetaminophen exhibited serious drawbacks, including liver damage, and is less effective in reducing inflammation compared to NSAIDs [33].
Neurotransmitter modulating agents, including amitriptyline and duloxetine, are also used in pain management. In neuropathic pain, anticonvulsants, including especially gabapentin and pregabalin, are typically first-line therapies [34]. For both these classes of drugs, side effects can include dizziness, drowsiness and weight gain [35]. In spite of these numerous therapies, the numerous risks and limited efficacy leave a tremendous unmet need for novel analgesic therapies [36].

2. Role of miRs in Pain

miRs are involved in the post-transcriptional regulation of genes [37] by binding to complementary sequences in target mRNAs. miR binding either inhibits translation or promotes mRNA degradation. Recent studies have shown that miRs modulate pain perception by regulating the expression of target genes involved in pain signaling pathways, including in nociceptors and spinal interneurons [38].

2.1. miR-124a

miR-124a is a key player in the spinal cord [39]. In a subset of spinal neurons in animal models, intraplantar formalin injection causes a considerable drop in miR-124a levels [40,41]. Moreover, miR-124 injection in the spinal cord decreases second-stage formalin-induced behaviors [38]. An independent study has shown that miR-124 directly regulates the expression of methyl-CpG binding protein 2 (MeCP2) [42]. MeCP2 is highly expressed in mature nerve cells and works as a transcriptional regulator of additional inflammatory pain genes [43,44].
Furthermore, it has been shown that the brain-derived neurotrophic factor (BDNF) protein, which is also well known as an established target gene of MECP2 that promotes nociceptive transmission in the dorsal horn, is downregulated by miR-124. Consequently, neuropathic pain following spinal cord damage may be influenced by miR-124 downregulation. Conversely, it has been demonstrated that miR-124 enhances microglia quiescence [45].
miR-124a has been shown in another investigation to suppress the expression of transient receptor potential vanilloid 1 (TRPV1), a crucial receptor involved in pain perception [46]. TRPV1 is a neuronal cation channel associated with sensitization to noxious stimuli, contributing to pain and inflammation [47]. Another mouse study found that glial released inflammatory factors during the central sensitization phase of neuropathic pain were linked to the shift in TRPV1 expression in adult mice’s cortical neurons. In this study, neuropathic pain was constitutively suppressed during early life by anti-inflammatory neuroimmune regulation [48]. In a second study, blood samples from patients with diverse neuropathic pain syndromes, such as polyneuropathy, postherpetic neuralgia, and trigeminal neuralgia, were found to express certain miRs, such as miR-124a and -155, differently in comparison to healthy persons [49]. Histone deacetylase sirtuin1 (SIRT1), a negative regulator of Foxp3, a master regulator of the formation of Tregs (regulatory T cells), is directly repressed by miR-124a and -155. Increased levels of miR-124a and miR-155 promote the differentiation of Tregs, which have an anti-inflammatory role, in neuropathic pain [50]. Additionally, the cerebral fluid of cancer patients who experienced pain showed downregulation of miR-124, indicating that miR-124 may be a useful analgesic medication for treating cancer pain sufferers [51].

2.2. miR-103

miR-103 is the first miR characterized in neuropathic pain [49]. In these reports, miR-103 functionally modulates nociceptors via transcripts, including calcium transients and their subunits (Cav1.2-α1, -α2δ1 and -β1) [52,53]. In neuropathic pain models, spinal dorsal horn neurons exhibit downregulation of miR-103, indicating that miR-103 may be a potential target for pain [54]. miR-103 has also been implicated in the regulation of the mu-opioid receptors (MORs). Inhibition of miR-103 led to increased MOR expression and enhanced the analgesic effects of opioids [55,56].

2.3. hsa-mir-548

The human miR-548 family contains 69 members located in almost all chromosomes [57]. A functional enrichment analysis of miR-548 gene family targets implicates a range of biological processes and diseases [58]. It has been reported that hsa-miR-548 expression is linked with the impaired interferon (IFN) signaling pathway in chronic hepatitis B (CHB) [59,60]. During the immunological activation (IA) phase of CHB, expression of hsa-miR-548-5p increases [59]. miR-548 may downregulate host antiviral response via direct targeting of IFN-λ1 [61], but miR-548 has been implicated in Wnt, MAPK, and TGF-β pathways as well [62]. According to another study, the 3′UTR of IFN-λ1 is the target of the miR-548 family members, which include miR-548b-5p, miR-548c-5p, miR-548i, miR-548j, and miR-548n. MiR-548 mimics inhibited IFN-λ1 expression, and, by controlling the host’s antiviral response, miR-548 could potentially be a treatment option for viruses [63].

2.4. miR-143

miR-143 has been implicated in the development and function of pain-sensing neurons in mice [64]. miR-143 is significantly downregulated (13.7 times) in tumors compared to the matched control that examined the function of miRs in pain and inflammation. [65]. Abnormal expression in miR-143 has been found in rheumatoid arthritis (RA) patients [66,67]. Fibromyalgia patients had decreased expression of hsa-miR-143-3p [68]; miR-143 also degrades COX-2 mRNA and regulates pain mediator production in pancreatic cancer [69]. After receiving a full Freund’s adjuvant (CFA) injection, ipsilateral Dorsal Root Ganglions (DRGs) showed noticeably reduced miR-143 expression levels [64]. Moreover, a bioinformatics analysis shows that around 1305 genes are associated with miR-143, most importantly mitogen-activated protein kinases MAPK7, MAPK3, COX2, matrix metalloproteinase, and TNF, all prominent target genes. This shows the significance of miR-143 in pain because all these target proteins are directly linked with various pain signaling pathways and the expression of miR-143 may control the expression of these target proteins [70].

2.5. miR-146a

miR-146a has been found upregulated in response to inflammatory stimuli and implicated in the negative feedback regulation of the inflammatory response [71,72,73]. miR-146a inhibits the production of cytokines and pain perception by targeting the mRNA of multiple pro-inflammatory genes, such as interleukin 1 receptor-associated kinase 1 (IRAK1) and tumor necrosis factor receptor-associated factor 6 (TRAF6) [74,75]. It has been proposed that the expression of Toll-like receptors (TLR) and nuclear factor kappa B (NF-κB) influences osteoarthritis pain [76]. These inflammatory response genes are regulated through the binding of mir-146 to the 3′-UTR (the three prime untranslated regions), inhibiting protein translation and reducing mRNA stability [77]. Therefore, miR-146 may act to reduce osteoarthritis pain driven by inflammatory cytokines [77,78]. In other pain models, a miR-146a-5p mimic inhibits TRAF6/c-Jun N-terminal kinases (JNK)/chemokine (C-C motif) ligand 2 (CCL2) signaling in astrocytes. miR-146a-5p inhibits spinal nerve ligation (SNL)-induced mechanical allodynia and spinal TRAF6 expression [79].

2.6. miR-let-7b

miR-let-7b is mainly involved in the transmission of pain signals in human spinal cord neurons [80]. In another study, the functional role of extracellular miRs in interaction with various receptors such as Toll-like receptor 7 (TLR7) and Transient receptor potential ankyrin 1 (TRPA1) to elicit pain, has been explored [81]. In DRG neurons, miR-let-7b causes fast inward currents and action potentials [80]. The most significant effect of let-7b is that it causes TLR7/TRPA1-dependent single-channel activity in HEK293 cells that overexpress TLR7/TRPA1 [82,83]. Let-7b injected in an intraplantar manner causes quick, spontaneous pain through the TLR7 and TRPA1 receptors. Ultimately, neuronal stimulation of DRG neurons can release let-7b and let-7b inhibitor diminishes formalin-induced TRPA1 currents and spontaneous pain [84]. Furthermore, it has been discovered that miR-let-7d controls the expression of the chemokine receptors CCR2 and CX3CR1, respectively, which are connected to pain and inflammation [85].

2.7. miR-21

miR-21 is linked with the activation of immune cells and regulation of both neuropathic pain and barrier disruption [86]. In peripheral neuropathy models, miR-21-5p is upregulated in DRG neurons [87]. It has also been observed that miR-21 levels increase in painful polyneuropathy both systemically in white blood cells and locally in sural-nerve biopsies [88]. miR-21-5p is believed to target a range of essential pain genes, including Matrix metalloproteinases (MMPs), phosphatase and tensin homolog (PTEN), and reversion-inducing cysteine-rich protein with Kazal motifs (RECKs) [89]. MMPs like MMP9 have a role in both barrier breakdown and the start of neuropathic pain [90]. Moreover, miR-21 may positively regulate transforming growth factor beta (TGF-β), an anti-inflammatory cytokine involved in barrier impairment and implicated in neuropathic pain [91,92,93]. TGF-β1 has been proposed as a critical mediator of nociception by inhibiting the neuroimmune feedback of both neurons and glia and by promoting the expression of endogenous opioids within the spinal cord [94].

2.8. miR-30c

miR-30c has been implicated in the neuropathic pain in mice [95]. Several pseudo receptors were found to be lacking in mice models with antiallodynic effects, including bone morphogenetic protein (BMP), activin membrane-bound inhibitor (BAMBI), and transforming growth factor-β (TGF-β) [96]. It has been observed that miR-30c is upregulated in the spinal cord, DRG, cerebrospinal fluid (CSF), and plasma, and miR-30c-5p expression positively correlates with allodynia [97]. In patients suffering from leg-ischemia-driven neuropathic pain, miR-30c-5p expression was increased in plasma and CSF compared to control patients without pain [97].
The miRs directly linked with different targets through differential expression in various tissues have been listed in Table 1.

3. miRs and Pain Signaling Pathways

Pain is primarily caused by four fundamental mechanistic processes: transduction, transmission, modulation, and perception [115]. Transduction is a process in which tissue-damaging stimuli activate nerve endings [116], and transmission is a transfer of a signal from the site of tissue injury to the brain regions that enable perception [117]. A brain function known as modulation works to selectively lower transmission system activity [118]. Perception is the subjective understanding governed by sensory signals [119]. Pain signaling pathways include ion channels, receptors, and intracellular signaling molecules that govern pain sensation with respect to causative factors. Certain miRs can regulate the chain reaction of signaling pathways. For instance, it has been demonstrated that miR-124 controls the expression of the N-methyl-D-aspartate (NMDA) receptor, which is essential for the emergence of chronic pain. The NMDA receptor, which is found in the spinal cord, is a therapeutic target for the treatment of neuropathic pain and plays a crucial role in nociceptive transmission and synaptic plasticity [120]. In some cases of neuropathic pain, it has been demonstrated that increased NMDA receptor activity leads to central sensitization. Thus, in animal models of neuropathic pain brought on by nerve damage, NMDA receptor antagonists can lessen hyperalgesia and allodynia [120]. Additionally, both miR-128a and miR-146a have also been involved in the regulation of voltage-gated sodium channel (Nav1.7) expression and intricate in pain signaling [121]. miRs can also regulate the expression of cytokines and chemokines that are involved in inflammation and pain. For instance, miR-146a has been shown to negatively regulate the expression of interleukin-1 receptor-associated kinase 1 (IRAK1), which plays a central role in the production of pro-inflammatory cytokines [122]. The overall scheme of this mechanism is shown in Figure 1.

3.1. miRs and Pain Mechanism

Many molecular mechanisms of pain have been associated with miRs, including brain plasticity and neurogenesis, nociceptor excitability, and chronic pain disorders [123,124]. Since brain-derived neurotrophic factor protein (BDNF) has been shown to modulate nociception, it is plausible that upregulating miR-132 in response to BDNF could enhance dendritic morphogenesis and arborization in the nociceptive pathways, thereby increasing the transmission of pain signals [82,125]. In another mouse study, it was observed that miR-29a/b plays a significant role in dendritic spine remodeling and synaptic plasticity [126]. Adult mice were given psychoactive substances like nicotine or cocaine in this experiment, and the outcomes demonstrated that miR-29a/b was consistently overexpressed in different brain areas. Furthermore, miR-29a/b was found to target the protein Arpc3, which is involved in actin branching during dendritic spine formation [127,128].
The therapeutic role of miRs in chronic pain has been revealed in earlier studies [40,129,130]. Numerous miRs, including miR-10a, miR-29a, miR-98, miR-99a, miR-124a, miR-134, and miR-183, have been significantly downregulated in the neurons of the ipsilateral trigeminal ganglion in an inflammatory rat model of CFA-induced muscular pain [124,131]. It was concluded that this downregulation of miRs increased the expression of multiple “pain-related” proteins, facilitating the onset of inflammation and allodynia [124]. On the other hand, mice with carrageenan-induced facial inflammation are shown to have an upregulated pair of miRNAs (miRNA-155/-223) in the pre-frontal cortex. Additionally, in this inflammatory model, an inverse association has been identified between miR-155 and the transcription factor, CCAAT/enhancer binding protein β [82]. In another mouse peritonitis model with acute inflammation, the enhanced levels of miR-21, miR-146b, miR-208a, miR-203, miR-142, miR-302d, and miR-219 were counter regulated by using resolvin D1 (RvD1; C22H32O5), a lipid mediator derived from docosahexaenoic acid that possesses anti-inflammatory and neuroprotective properties [132]. This study clearly shows that specific miRs act as post-transcriptional regulators of the inflammatory pain condition [133].

3.2. miRs in Visceral Pain

Visceral pain is the condition in which internal organs become inflamed, diseased, damaged, or injured, such as bladder pain, endometriosis, irritable bowel syndrome, and prostate pain [134,135]. Published data indicates that greater expression of miR-29a in colon tissues and blood micro-vesicles of persons with irritable bowel syndrome (IBS) is positively connected with improved intestinal membrane permeability [136]. It is well known that miR-29a inhibits the production of glutamate ammonia ligase (GLUL) by interacting with complementary binding sites at the 3′UTR. This increases intestinal permeability and eventually results in persistent visceral pain [136]. Increased expression levels of miR-328, miR-320, miR-449b, and miR-500 have been linked to the downregulation of neurokinin-1 (NK1) receptors in bladder biopsies in a study of individuals with bladder pain syndrome (BPS) [109]. Other published data indicate the possible relation of miR-199a-5p with urothelial permeability in bladder pain syndrome [137]. Another possible role of miRs has been observed in patients with painful endometriosis [138,139]. Additional studies have shown that miR-181, miR-216, and miR-203 are the known targets in regard to the GABAAα-1 gene by using a luciferase reporter assay [113,140]. The overall mechanism is depicted in Figure 2.

3.3. Polymorphisms of miRs and Pain

The change in the miRNA expression leads to genetic mutations which are associated with the function of particular genes [141]. However, there are still many studies that explore the significant impact of polymorphisms in miR genes and/or their targets, which are closely related to nociception. The single nucleotide polymorphisms (SNPs) in miRNAs are associated with the pathogenesis of different diseases. Furthermore, SNPs may also affect the functions of genes through the alteration in miRNA expression. Analogously, SNPs in the miRs’ target sites have the ability to modify or disrupt the miRs’ already-existing binding sites as well as to generate new ones [142]. It has been noted that rheumatoid arthritis and miR polymorphism are directly related. The significant correlations between psoriatic arthritis (PsA) and miR-146a with rs2910164 (G/C) have been observed in both heterozygous and dominant models. Additionally, the heterozygous model demonstrated a strong correlation between ankylosing spondylitis (AS) and the miR-146a rs2910164 (G/C) polymorphism [143]. The overall depiction is shown in Figure 3.

4. miRNAs as Therapeutic Targets

Multiple direct and indirect approaches have been utilized to target miRs expression in pain [144,145]. The direct strategies directly use the oligonucleotides or virus-based constructs to either block the overexpression of miRNAs or to substitute for the loss of expression of miRNAs [146]. The indirect strategies are linked with the utilization of drugs to modulate miRNA expression by regulating their transcription and processing [147]. Nucleotide sequences mimic mature or pre-miRNAs to overexpress miRNAs, while anti-sense oligos (anti-miRNAs), miRNA sponges, or miRNA decoys are typically used to limit miRNA expression [82]. The most important task is to deliver miRNAs into the central nervous system (CNS), which is the blood–brain barrier (BBB). However, there are a few techniques for CNS gene silencing, including germline transfection, lentiviral vector-mediated administration, and stereotaxic administration [148,149]. An alternative method is based on immunoliposomes, which are a combination of liposomes, tiny hairpin RNA expression plasmids, and a monoclonal antibody directed toward a specific receptor. Immunoliposome-based targeted delivery makes it easier for conjugated RNA to pass across the blood–brain barrier and enter particular brain areas [150,151].

miRs as Prognostic Biomarkers of Pain

Biomarkers are measurable characteristics of a disease state and are crucial components for more precisely delivering tailored therapy. Furthermore, dysregulated miR expression in bodily fluids is being investigated as a non-invasive clinical diagnostic for pain [152,153,154,155]. Multiple studies have explored changes in miRNA expression in the inflammatory pain-affected tissue samples in rodent models [88,155,156]. It has also been observed that miRs such as miR-103, miR-124a, miR-203, and miR-7a have a direct influence on the management of pain [52,157].
The association among pain, plasma cytokines, and miRs in blood has been detected [158]. In this study, miRs expression level investigations were used to examine twenty-two controls, thirty-three individuals with persistent pelvic discomfort, and twenty-three individuals with Vertebrobasilar dolichoectasia (VBD), a disorder marked by elongation, tortuosity, and ectasia of the basilar artery [158]. VBS may manifest clinically by compression of the cranial nerves, ischemic symptoms, or intracranial bleeding and irritable bowel syndrome (IBS). It has been noted that miR dysregulation in VBD is expected to impact estrogen-relevant pathways, which may play a role in pelvic-localized pain. Moreover, changes in muscle, neuron, and glial cell function have been linked to miRNA abnormalities in both VBD and IBS patients, which in turn affect pain [158]. miR-132 modulates cholinergic signaling and inflammation in human inflammatory bowel disease. Compared to healthy controls, IBD patients had higher levels of miR-132 and, accordingly, decreased acetylcholinesterase (AChE) activity. These findings support the idea that inflammation plays a homeostatic function in cholinergic transmission [159,160].

5. miRs-Gene Association Networks in Pain

Numerous miRs have been found to be connected to genes involved in chronic pain. Dysregulation of these miRs can affect the expression of genes related to neuroinflammation, neural plasticity, and pain signaling. Table 2 lists some known instances of miRs associated with genes linked to chronic pain.

6. Conclusions

Our review highlighted basic mechanisms of miR regulation of pain through a range of molecular targets expressed primarily in neurons. The expression levels of miRs are characteristic of neuronal encoding of different pain states. miRs play an essential role in the molecular etiology of chronic pain and offer potential as novel therapeutic targets.

Author Contributions

Conceptualization, M.H. and A.K.; data curation, S.S. and M.Y.; writing—original draft preparation, M.H. and W.C.; writing—review and editing, A.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by NIH grant numbers R01GM127701 and R01HG012117.

Acknowledgments

A.K. acknowledges financial support from NIH Grants R01GM127701 and R01HG012117. M.H. acknowledges The Ohio State University for providing the “President’s Postdoctoral Scholars Program (PPSP)” award and financial support to complete this study.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Mendell, L.M.; Albers, K.M.; Davis, B.M. Neurotrophins, nociceptors, and pain. Microsc. Res. Tech. 1999, 45, 252–261. [Google Scholar] [CrossRef]
  2. Wang, M.; Thyagarajan, B. Pain pathways and potential new targets for pain relief. Biotechnol. Appl. Biochem. 2022, 69, 110–123. [Google Scholar] [CrossRef] [PubMed]
  3. Pattison, L.A.; Callejo, G.; St John Smith, E. Evolution of acid nociception: Ion channels and receptors for detecting acid. Philos. Trans. R. Soc. B 2019, 374, 20190291. [Google Scholar] [CrossRef] [PubMed]
  4. Steeds, C.E. The anatomy and physiology of pain. Surgery 2009, 27, 507–511. [Google Scholar]
  5. D’Mello, R.; Dickenson, A.H. Spinal cord mechanisms of pain. Br. J. Anaesth. 2008, 101, 8–16. [Google Scholar] [CrossRef]
  6. Melzack, R. Pain: Past, present and future. Can. J. Exp. Psychol./Rev. Can. Psychol. Expérimentale 1993, 47, 615. [Google Scholar] [CrossRef] [PubMed]
  7. Świeboda, P.; Filip, R.; Prystupa, A.; Drozd, M. Assessment of pain: Types, mechanism and treatment. Pain 2013, 2, 2–7. [Google Scholar]
  8. Tracey, W.D. Nociception. Curr. Biol. 2017, 27, R129–R133. [Google Scholar] [CrossRef] [PubMed]
  9. Tigerholm, J.; Poulsen, A.H.; Andersen, O.K.; Mørch, C.D. From perception threshold to ion channels—A computational study. Biophys. J. 2019, 117, 281–295. [Google Scholar] [CrossRef] [PubMed]
  10. Schliessbach, J.; Maurer, K. Pharmacology of pain transmission and modulation. Pain Med. Essent. Rev. 2017, 7–9. [Google Scholar]
  11. Zhou, H.-Y.; Zhang, H.-M.; Chen, S.-R.; Pan, H.-L. Increased C-fiber nociceptive input potentiates inhibitory glycinergic transmission in the spinal dorsal horn. J. Pharmacol. Exp. Ther. 2008, 324, 1000–1010. [Google Scholar] [CrossRef] [PubMed]
  12. Tsuda, M.; Koga, K.; Chen, T.; Zhuo, M. Neuronal and microglial mechanisms for neuropathic pain in the spinal dorsal horn and anterior cingulate cortex. J. Neurochem. 2017, 141, 486–498. [Google Scholar] [CrossRef] [PubMed]
  13. Cohen, S.P.; Mao, J. Neuropathic pain: Mechanisms and their clinical implications. Bmj 2014, 348, f7656. [Google Scholar] [CrossRef] [PubMed]
  14. Katz, J.; Rosenbloom, B.N.; Fashler, S. Chronic pain, psychopathology, and DSM-5 somatic symptom disorder. Can. J. Psychiatry 2015, 60, 160–167. [Google Scholar] [CrossRef] [PubMed]
  15. Jensen, T.S.; Finnerup, N.B. Allodynia and hyperalgesia in neuropathic pain: Clinical manifestations and mechanisms. Lancet Neurol. 2014, 13, 924–935. [Google Scholar] [CrossRef] [PubMed]
  16. Jensen, T.S.; Gottrup, H.; Sindrup, S.H.; Bach, F.W. The clinical picture of neuropathic pain. Eur. J. Pharmacol. 2001, 429, 1–11. [Google Scholar] [CrossRef] [PubMed]
  17. Wahid, F.; Shehzad, A.; Khan, T.; Kim, Y.Y. MicroRNAs: Synthesis, mechanism, function, and recent clinical trials. Biochim. Et Biophys. Acta (BBA)-Mol. Cell Res. 2010, 1803, 1231–1243. [Google Scholar] [CrossRef] [PubMed]
  18. Ying, S.-Y.; Chang, D.C.; Lin, S.-L. The microRNA (miRNA): Overview of the RNA genes that modulate gene function. Mol. Biotechnol. 2008, 38, 257–268. [Google Scholar] [CrossRef] [PubMed]
  19. Kim, V.N. MicroRNA biogenesis: Coordinated cropping and dicing. Nature Rev. Mol. Cell Biol. 2005, 6, 376–385. [Google Scholar] [CrossRef] [PubMed]
  20. Di Leva, G.; Calin, G.A.; Croce, C.M. MicroRNAs: Fundamental facts and involvement in human diseases. Birth Defects Res. Part C Embryo Today Rev. 2006, 78, 180–189. [Google Scholar] [CrossRef] [PubMed]
  21. Yaksh, T.L.; Wallace, M.S. Opioids, analgesia, and pain management. In Goodman and Gilman’s the Pharmacological Basis of Therapeutics; Academia: San Francisco, CA, USA, 2011; pp. 481–526. [Google Scholar]
  22. Barkin, R.L. Acetaminophen, aspirin, or ibuprofen in combination analgesic products. Am. J. Ther. 2001, 8, 433–442. [Google Scholar] [CrossRef] [PubMed]
  23. Green, G.A. Understanding NSAIDs: From aspirin to COX-2. Clin. Cornerstone 2001, 3, 50–59. [Google Scholar] [CrossRef] [PubMed]
  24. Russell, R. Non-steroidal anti-inflammatory drugs and gastrointestinal damage—Problems and solutions. Postgrad. Med. J. 2001, 77, 82–88. [Google Scholar] [CrossRef] [PubMed]
  25. Varas-Lorenzo, C.; Riera-Guardia, N.; Calingaert, B.; Castellsague, J.; Pariente, A.; Scotti, L.; Sturkenboom, M.; Perez-Gutthann, S. Stroke risk and NSAIDs: A systematic review of observational studies. Pharmacoepidemiol. Drug Saf. 2011, 20, 1225–1236. [Google Scholar] [CrossRef] [PubMed]
  26. Ejaz, P.; Bhojani, K.; Joshi, V. NSAIDs and kidney. Japi 2004, 52, 371. [Google Scholar]
  27. Prados-Torres, A.; Calderón-Larrañaga, A.; Hancco-Saavedra, J.; Poblador-Plou, B.; van den Akker, M. Multimorbidity patterns: A systematic review. J. Clin. Epidemiol. 2014, 67, 254–266. [Google Scholar] [CrossRef] [PubMed]
  28. Owusu Obeng, A.; Hamadeh, I.; Smith, M. Review of opioid pharmacogenetics and considerations for pain management. Pharmacother. J. Hum. Pharmacol. Drug Ther. 2017, 37, 1105–1121. [Google Scholar] [CrossRef] [PubMed]
  29. Mercadante, S. Opioid analgesics adverse effects: The other side of the coin. Curr. Pharm. Des. 2019, 25, 3197–3202. [Google Scholar] [CrossRef] [PubMed]
  30. Brown, R.T.; Zuelsdorff, M.; Fleming, M. Adverse effects and cognitive function among primary care patients taking opioids for chronic nonmalignant pain. J. Opioid Manag. 2006, 2, 137–146. [Google Scholar] [CrossRef] [PubMed]
  31. Becker, W.C.; Sullivan, L.E.; Tetrault, J.M.; Desai, R.A.; Fiellin, D.A. Non-medical use, abuse and dependence on prescription opioids among US adults: Psychiatric, medical and substance use correlates. Drug Alcohol Depend. 2008, 94, 38–47. [Google Scholar] [CrossRef]
  32. Ayoub, S.S. Paracetamol (acetaminophen): A familiar drug with an unexplained mechanism of action. Temperature 2021, 8, 351–371. [Google Scholar] [CrossRef] [PubMed]
  33. Gupta, A.; Jakobsson, J. Acetaminophen, nonsteroidal anti-inflammatory drugs, and cyclooxygenase-2 selective inhibitors: An update. Plast. Reconstr. Surg. 2014, 134, 24S–31S. [Google Scholar] [CrossRef] [PubMed]
  34. Söderpalm, B. Anticonvulsants: Aspects of their mechanisms of action. Eur. J. Pain 2002, 6, 3–9. [Google Scholar] [CrossRef] [PubMed]
  35. Swann, A.C. Major system toxicities and side effects of anticonvulsants. J. Clin. Psychiatry 2001, 62, 16–21. [Google Scholar] [PubMed]
  36. Nagakura, Y. The need for fundamental reforms in the pain research field to develop innovative drugs. Expert Opin. Drug Discov. 2017, 12, 39–46. [Google Scholar] [CrossRef] [PubMed]
  37. Dykes, I.M.; Emanueli, C. Transcriptional and post-transcriptional gene regulation by long non-coding RNA. Genom. Proteom. Bioinform. 2017, 15, 177–186. [Google Scholar] [CrossRef]
  38. Dai, Z.; Chu, H.; Ma, J.; Yan, Y.; Zhang, X.; Liang, Y. The regulatory mechanisms and therapeutic potential of microRNAs: From chronic pain to morphine tolerance. Front. Mol. Neurosci. 2018, 11, 80. [Google Scholar] [CrossRef] [PubMed]
  39. Nakanishi, K.; Nakasa, T.; Tanaka, N.; Ishikawa, M.; Yamada, K.; Yamasaki, K.; Kamei, N.; Izumi, B.; Adachi, N.; Miyaki, S. Responses of microRNAs 124a and 223 following spinal cord injury in mice. Spinal Cord 2010, 48, 192–196. [Google Scholar] [CrossRef] [PubMed]
  40. Sakai, A.; Suzuki, H. Emerging roles of microRNAs in chronic pain. Neurochem. Int. 2014, 77, 58–67. [Google Scholar] [CrossRef] [PubMed]
  41. Kynast, K.L.; Russe, O.Q.; Möser, C.V.; Geisslinger, G.; Niederberger, E. Modulation of central nervous system–specific microRNA-124a alters the inflammatory response in the formalin test in mice. Pain 2013, 154, 368–376. [Google Scholar] [CrossRef]
  42. Amir, R.E.; Van den Veyver, I.B.; Wan, M.; Tran, C.Q.; Francke, U.; Zoghbi, H.Y. Rett syndrome is caused by mutations in X-linked MECP2, encoding methyl-CpG-binding protein 2. Nat. Genet. 1999, 23, 185–188. [Google Scholar] [CrossRef] [PubMed]
  43. Manners, M.T.; Ertel, A.; Tian, Y.; Ajit, S.K. Genome-wide redistribution of MeCP2 in dorsal root ganglia after peripheral nerve injury. Epigenetics Chromatin 2016, 9, 1–15. [Google Scholar] [CrossRef] [PubMed]
  44. Géranton, S.M.; Morenilla-Palao, C.; Hunt, S.P. A role for transcriptional repressor methyl-CpG-binding protein 2 and plasticity-related gene serum-and glucocorticoid-inducible kinase 1 in the induction of inflammatory pain states. J. Neurosci. 2007, 27, 6163–6173. [Google Scholar] [CrossRef] [PubMed]
  45. Ponomarev, E.D.; Veremeyko, T.; Barteneva, N.; Krichevsky, A.M.; Weiner, H.L. MicroRNA-124 promotes microglia quiescence and suppresses EAE by deactivating macrophages via the C/EBP-α–PU. 1 pathway. Nat. Med. 2011, 17, 64–70. [Google Scholar] [CrossRef] [PubMed]
  46. Hu, C.; He, M.; Xu, Q.; Tian, W. Advances with Non-Coding RNAs in Neuropathic Pain. Front. Neurosci. 2021, 15, 1574. [Google Scholar] [CrossRef] [PubMed]
  47. Laing, R.J.; Dhaka, A. ThermoTRPs and pain. Neuroscientist 2016, 22, 171–187. [Google Scholar] [CrossRef] [PubMed]
  48. Marrone, M.C.; Morabito, A.; Giustizieri, M.; Chiurchiu, V.; Leuti, A.; Mattioli, M.; Marinelli, S.; Riganti, L.; Lombardi, M.; Murana, E. TRPV1 channels are critical brain inflammation detectors and neuropathic pain biomarkers in mice. Nat. Commun. 2017, 8, 15292. [Google Scholar] [CrossRef] [PubMed]
  49. Morchio, M.; Sher, E.; Collier, D.A.; Lambert, D.W.; Boissonade, F.M. The Role of miRNAs in Neuropathic Pain. Biomedicines 2023, 11, 775. [Google Scholar] [CrossRef] [PubMed]
  50. Heyn, J.; Luchting, B.; Hinske, L.C.; Hübner, M.; Azad, S.C.; Kreth, S. miR-124a and miR-155 enhance differentiation of regulatory T cells in patients with neuropathic pain. J. Neuroinflamm. 2016, 13, 1–11. [Google Scholar] [CrossRef] [PubMed]
  51. Elramah, S.; López-González, M.J.; Bastide, M.; Dixmérias, F.; Roca-Lapirot, O.; Wielanek-Bachelet, A.-C.; Vital, A.; Leste-Lasserre, T.; Brochard, A.; Landry, M. Spinal miRNA-124 regulates synaptopodin and nociception in an animal model of bone cancer pain. Sci. Rep. 2017, 7, 10949. [Google Scholar] [CrossRef]
  52. Favereaux, A.; Thoumine, O.; Bouali-Benazzouz, R.; Roques, V.; Papon, M.A.; Salam, S.A.; Drutel, G.; Léger, C.; Calas, A.; Nagy, F. Bidirectional integrative regulation of Cav1. 2 calcium channel by microRNA miR-103: Role in pain. EMBO J. 2011, 30, 3830–3841. [Google Scholar] [CrossRef] [PubMed]
  53. Fossat, P.; Dobremez, E.; Bouali-Benazzouz, R.; Favereaux, A.; Bertrand, S.S.; Kilk, K.; Léger, C.; Cazalets, J.-R.; Langel, Ü.; Landry, M. Knockdown of L calcium channel subtypes: Differential effects in neuropathic pain. J. Neurosci. 2010, 30, 1073–1085. [Google Scholar] [CrossRef] [PubMed]
  54. Shi, G.; Shi, J.; Liu, K.; Liu, N.; Wang, Y.; Fu, Z.; Ding, J.; Jia, L.; Yuan, W. Increased miR-195 aggravates neuropathic pain by inhibiting autophagy following peripheral nerve injury. Glia 2013, 61, 504–512. [Google Scholar] [CrossRef] [PubMed]
  55. Lu, Z.; Xu, J.; Xu, M.; Pasternak, G.W.; Pan, Y.-X. Morphine regulates expression of μ-opioid receptor MOR-1A, an intron-retention carboxyl terminal splice variant of the μ-opioid receptor (OPRM1) gene via miR-103/miR-107. Mol. Pharmacol. 2014, 85, 368–380. [Google Scholar] [CrossRef] [PubMed]
  56. Lee, S. Pharmacological inhibition of voltage-gated Ca2+ channels for chronic pain relief. Curr. Neuropharmacol. 2013, 11, 606–620. [Google Scholar] [PubMed]
  57. Liang, T.; Guo, L.; Liu, C. Genome-wide analysis of mir-548 gene family reveals evolutionary and functional implications. BioMed Res. Int. 2012, 2012, 679563. [Google Scholar] [CrossRef] [PubMed]
  58. Calhoun, M.A.; Cui, Y.; Elliott, E.E.; Mo, X.; Otero, J.J.; Winter, J.O. MicroRNA-mRNA interactions at low levels of compressive solid stress implicate mir-548 in increased glioblastoma cell motility. Sci. Rep. 2020, 10, 311. [Google Scholar] [CrossRef] [PubMed]
  59. Li, N.; Chen, M.; Cheng, Q.; Zhu, M.; Shi, G. Identification of microRNA mediating type I interferon pathway in chronic hepatitis B patients. Chin. J. Viral Dis 2012, 2, 107–112. [Google Scholar]
  60. Xing, T.; Xu, H.; Yu, W.; Wang, B.; Zhang, J. Expression profile and clinical significance of miRNAs at different stages of chronic hepatitis B virus infection. Int. J. Clin. Exp. Med. 2015, 8, 5611. [Google Scholar] [PubMed]
  61. Ramos-Sanchez, E.M.; Reis, L.C.; Souza MD, A.; Muxel, S.M.; Santos, K.R.; Lagos, D.; Goto, H. miR-548d-3p is up-regulated in human visceral leishmaniasis and suppresses parasite growth in macrophages. Front. Cell. Infect. Microbiol. 2022, 12, 826039. [Google Scholar]
  62. Xing, T.-J.; Xu, H.-T.; Yu, W.-Q.; Wang, B.; Zhang, J. MiRNA-548ah, a potential molecule associated with transition from immune tolerance to immune activation of chronic hepatitis B. Int. J. Mol. Sci. 2014, 15, 14411–14426. [Google Scholar] [CrossRef] [PubMed]
  63. Li, Y.; Xie, J.; Xu, X.; Wang, J.; Ao, F.; Wan, Y.; Zhu, Y. MicroRNA-548 down-regulates host antiviral response via direct targeting of IFN-λ1. Protein Cell 2013, 4, 130–141. [Google Scholar] [CrossRef] [PubMed]
  64. Tam Tam, S.; Bastian, I.; Zhou, X.-F.; Vander Hoek, M.; Michael, M.Z.; Gibbins, I.L.; Haberberger, R. MicroRNA-143 expression in dorsal root ganglion neurons. Cell Tissue Res. 2011, 346, 163–173. [Google Scholar] [CrossRef] [PubMed]
  65. Lin, T.; Dong, W.; Huang, J.; Pan, Q.; Fan, X.; Zhang, C.; Huang, L. MicroRNA-143 as a tumor suppressor for bladder cancer. J. Urol. 2009, 181, 1372–1380. [Google Scholar] [CrossRef] [PubMed]
  66. Zhu, Y.; Zhou, L.; Peng, X.; Geng, S.; Zhou, X. Study of Qingluo Tongbi Compound treating rheumatoid arthritis based on miRNA network. Chin. J. Immunol. 2016, 12, 495–499. [Google Scholar]
  67. Yuan, C.; Geng, S.; Zhu, Y.; Peng, X.; Zhou, X.; Zhou, L. Comparison of abnormal expression of miRNAs in peripheral blood of rheumatoid arthritis patients and osteoclasts in rat and analysis of any miRNAs. Chin. J. Immunol. 2017, 12, 715–720. [Google Scholar]
  68. Cerdá-Olmedo, G.; Mena-Durán, A.V.; Monsalve, V.; Oltra, E. Identification of a microRNA signature for the diagnosis of fibromyalgia. PLoS ONE 2015, 10, e0121903. [Google Scholar] [CrossRef] [PubMed]
  69. Pham, H.; Rodriguez, C.E.; Donald, G.W.; Hertzer, K.M.; Jung, X.S.; Chang, H.-H.; Moro, A.; Reber, H.A.; Hines, O.J.; Eibl, G. miR-143 decreases COX-2 mRNA stability and expression in pancreatic cancer cells. Biochem. Biophys. Res. Commun. 2013, 439, 6–11. [Google Scholar] [CrossRef] [PubMed]
  70. Zhou, L.L.; Zhu, Y.M.; Qian, F.Y.; Yuan, C.C.; Yuan, D.P.; Zhou, X.P. MicroRNA-143-3p contributes to the regulation of pain responses in collagen-induced arthritis. Mol. Med. Rep. 2018, 18, 3219–3228. [Google Scholar] [CrossRef] [PubMed]
  71. Iyer, A.; Zurolo, E.; Prabowo, A.; Fluiter, K.; Spliet, W.G.; van Rijen, P.C.; Gorter, J.A.; Aronica, E. MicroRNA-146a: A key regulator of astrocyte-mediated inflammatory response. PLoS ONE 2012, e44789. [Google Scholar] [CrossRef] [PubMed]
  72. Meisgen, F.; Landén, N.X.; Wang, A.; Réthi, B.; Bouez, C.; Zuccolo, M.; Gueniche, A.; Ståhle, M.; Sonkoly, E.; Breton, L. MiR-146a negatively regulates TLR2-induced inflammatory responses in keratinocytes. J. Investig. Dermatol. 2014, 134, 1931–1940. [Google Scholar] [CrossRef] [PubMed]
  73. Roos, J.; Enlund, E.; Funcke, J.-B.; Tews, D.; Holzmann, K.; Debatin, K.-M.; Wabitsch, M.; Fischer-Posovszky, P. miR-146a-mediated suppression of the inflammatory response in human adipocytes. Sci. Rep. 2016, 6, 38339. [Google Scholar] [CrossRef] [PubMed]
  74. Li, X.; Gibson, G.; Kim, J.-S.; Kroin, J.; Xu, S.; Van Wijnen, A.J.; Im, H.-J. MicroRNA-146a is linked to pain-related pathophysiology of osteoarthritis. Gene 2011, 480, 34–41. [Google Scholar] [CrossRef] [PubMed]
  75. Mortazavi-Jahromi, S.S.; Ahmadzadeh, A.; Rezaieyazdi, Z.; Aslani, M.; Omidian, S.; Mirshafiey, A. The role of β-d-mannuronic acid, as a new non-steroidal anti-inflammatory drug on expression of miR-146a, IRAK1, TRAF6, NF-κB and pro-inflammatory cytokines following a clinical trial in rheumatoid arthritis patients. Immunopharmacol. Immunotoxicol. 2020, 42, 228–236. [Google Scholar] [CrossRef] [PubMed]
  76. Liu-Bryan, R.; Terkeltaub, R. Chondrocyte innate immune myeloid differentiation factor 88–dependent signaling drives procatabolic effects of the endogenous toll-like receptor 2/toll-like receptor 4 ligands low molecular weight hyaluronan and high mobility group box chromosomal protein 1 in mice. Arthritis Rheum. 2010, 62, 2004–2012. [Google Scholar]
  77. Taganov, K.D.; Boldin, M.P.; Chang, K.-J.; Baltimore, D. NF-κB-dependent induction of microRNA miR-146, an inhibitor targeted to signaling proteins of innate immune responses. Proc. Natl. Acad. Sci. USA 2006, 103, 12481–12486. [Google Scholar] [CrossRef] [PubMed]
  78. Bhaumik, D.; Scott, G.; Schokrpur, S.; Patil, C.; Campisi, J.; Benz, C. Expression of microRNA-146 suppresses NF-κB activity with reduction of metastatic potential in breast cancer cells. Oncogene 2008, 27, 5643–5647. [Google Scholar] [CrossRef]
  79. Lu, Y.; Cao, D.-L.; Jiang, B.-C.; Yang, T.; Gao, Y.-J. MicroRNA-146a-5p attenuates neuropathic pain via suppressing TRAF6 signaling in the spinal cord. Brain Behav. Immun. 2015, 49, 119–129. [Google Scholar] [CrossRef] [PubMed]
  80. Ji, R.-R. Neuroimmune interactions in itch: Do chronic itch, chronic pain, and chronic cough share similar mechanisms? Pulm. Pharmacol. Ther. 2015, 35, 81–86. [Google Scholar] [CrossRef] [PubMed]
  81. Santoni, G.; Cardinali, C.; Morelli, M.B.; Santoni, M.; Nabissi, M.; Amantini, C. Danger-and pathogen-associated molecular patterns recognition by pattern-recognition receptors and ion channels of the transient receptor potential family triggers the inflammasome activation in immune cells and sensory neurons. J. Neuroinflamm. 2015, 12, 1–10. [Google Scholar] [CrossRef]
  82. Zhang, J.; Banerjee, B. Role of microRNA in visceral pain. J. Neurogastroenterol. Motil. 2015, 21, 159. [Google Scholar] [CrossRef] [PubMed]
  83. Adlakha, Y.K. MicroRNAs: The tiny robust players unraveling the multifaceted channels of pain. Benth. Sci. 2016, 26, 126–160. [Google Scholar]
  84. Park, C.-K.; Xu, Z.-Z.; Berta, T.; Han, Q.; Chen, G.; Liu, X.-J.; Ji, R.-R. Extracellular microRNAs activate nociceptor neurons to elicit pain via TLR7 and TRPA1. Neuron 2014, 82, 47–54. [Google Scholar] [CrossRef] [PubMed]
  85. Buonfiglioli, A.; Efe, I.E.; Guneykaya, D.; Ivanov, A.; Huang, Y.; Orlowski, E.; Krüger, C.; Deisz, R.A.; Markovic, D.; Flüh, C. let-7 MicroRNAs regulate microglial function and suppress glioma growth through toll-like receptor 7. Cell Rep. 2019, 29, 3460–3471.e3467. [Google Scholar] [CrossRef] [PubMed]
  86. Zhong, L.; Xiao, W.; Wang, F.; Liu, J.; Zhi, L.-J. miR-21-5p inhibits neuropathic pain development via directly targeting C-C motif ligand 1 and tissue inhibitor of metalloproteinase-3. J. Cell. Biochem. 2019, 120, 16614–16623. [Google Scholar] [CrossRef] [PubMed]
  87. Sakai, A.; Suzuki, H. Nerve injury-induced upregulation of miR-21 in the primary sensory neurons contributes to neuropathic pain in rats. Biochem. Biophys. Res. Commun. 2013, 435, 176–181. [Google Scholar] [CrossRef] [PubMed]
  88. Leinders, M.; Üçeyler, N.; Thomann, A.; Sommer, C. Aberrant microRNA expression in patients with painful peripheral neuropathies. J. Neurol. Sci. 2017, 380, 242–249. [Google Scholar] [CrossRef] [PubMed]
  89. Kawasaki, Y.; Xu, Z.-Z.; Wang, X.; Park, J.Y.; Zhuang, Z.-Y.; Tan, P.-H.; Gao, Y.-J.; Roy, K.; Corfas, G.; Lo, E.H. Distinct roles of matrix metalloproteases in the early-and late-phase development of neuropathic pain. Nat. Med. 2008, 14, 331–336. [Google Scholar] [CrossRef] [PubMed]
  90. Hackel, D.; Krug, S.M.; Sauer, R.-S.; Mousa, S.A.; Böcker, A.; Pflücke, D.; Wrede, E.-J.; Kistner, K.; Hoffmann, T.; Niedermirtl, B. Transient opening of the perineurial barrier for analgesic drug delivery. Proc. Natl. Acad. Sci. USA 2012, 109, E2018–E2027. [Google Scholar] [CrossRef] [PubMed]
  91. Wang, K.; Pascal, L.E.; Li, F.; Chen, W.; Dhir, R.; Balasubramani, G.K.; DeFranco, D.B.; Yoshimura, N.; He, D.; Wang, Z. Tight junction protein claudin-1 is downregulated by TGF-β1 via MEK signaling in benign prostatic epithelial cells. Prostate 2020, 80, 1203–1215. [Google Scholar] [CrossRef] [PubMed]
  92. Kojima, T.; Takano, K.i.; Yamamoto, T.; Murata, M.; Son, S.; Imamura, M.; Yamaguchi, H.; Osanai, M.; Chiba, H.; Himi, T. Transforming growth factor-β induces epithelial to mesenchymal transition by down-regulation of claudin-1 expression and the fence function in adult rat hepatocytes. Liver Int. 2008, 28, 534–545. [Google Scholar] [CrossRef] [PubMed]
  93. Reinhold, A.K.; Krug, S.M.; Salvador, E.; Sauer, R.S.; Karl-Schöller, F.; Malcangio, M.; Sommer, C.; Rittner, H.L. MicroRNA-21-5p functions via RECK/MMP9 as a proalgesic regulator of the blood nerve barrier in nerve injury. Ann. N. Y. Acad. Sci. 2022, 1515, 184–195. [Google Scholar] [CrossRef] [PubMed]
  94. Lantero, A.; Tramullas, M.; Díaz, A.; Hurlé, M.A. Transforming growth factor-β in normal nociceptive processing and pathological pain models. Mol. Neurobiol. 2012, 45, 76–86. [Google Scholar] [CrossRef] [PubMed]
  95. Tramullas, M.; Francés, R.; de la Fuente, R.; Velategui, S.; Carcelén, M.; García, R.; Hurlé, M.A. MicroRNA-30c-5p modulates neuropathic pain in rodents. Sci. Transl. Med. 2018, 10, eaao6299. [Google Scholar] [CrossRef] [PubMed]
  96. Fidilio, A.; Grasso, M.; Turnaturi, R.; Caruso, G.; Spitale, F.M.; Vicario, N.; Parenti, R.; Spoto, S.; Musso, N.; Marrazzo, A. The multimodal MOPr/DOPr agonist LP2 reduces allodynia in chronic constriction injured rats by rescue of TGF-β1 signalling. Front. Pharmacol. 2021, 12, 749365. [Google Scholar] [CrossRef] [PubMed]
  97. Francés, R.; Mata-Garrido, J.; de la Fuente, R.; Carcelén, M.; Lafarga, M.; Berciano, M.T.; Tramullas, M. Identification of epigenetic interactions between MicroRNA-30c-5p and DNA methyltransferases in neuropathic pain. Int. J. Mol. Sci. 2022, 23, 13994. [Google Scholar] [CrossRef] [PubMed]
  98. Tao, Z.-Y.; Xue, Y.; Li, J.-F.; Traub, R.J.; Cao, D.-Y. Do microRNAs modulate visceral pain? BioMed Res. Int. 2018, 2018, 5406973. [Google Scholar] [CrossRef] [PubMed]
  99. Zhou, Q.; Yang, L.; Larson, S.; Basra, S.; Merwat, S.; Tan, A.; Croce, C.; Verne, G.N. Decreased miR-199 augments visceral pain in patients with IBS through translational upregulation of TRPV1. Gut 2015, 65, 797–805. [Google Scholar] [CrossRef] [PubMed]
  100. Liao, X.-J.; Mao, W.-M.; Wang, Q.; Yang, G.-G.; Wu, W.-J.; Shao, S.-X. MicroRNA-24 inhibits serotonin reuptake transporter expression and aggravates irritable bowel syndrome. Biochem. Biophys. Res. Commun. 2016, 469, 288–293. [Google Scholar] [CrossRef] [PubMed]
  101. Bradesi, S.; Karagiannides, I.; Bakirtzi, K.; Joshi, S.M.; Koukos, G.; Iliopoulos, D.; Pothoulakis, C.; Mayer, E.A. Identification of spinal cord MicroRNA and gene signatures in a model of chronic stress-induced visceral hyperalgesia in rat. PLoS ONE 2015, 10, e0130938. [Google Scholar] [CrossRef] [PubMed]
  102. Fourie, N.H.; Peace, R.M.; Abey, S.K.; Sherwin, L.B.; Rahim-Williams, B.; Smyser, P.A.; Wiley, J.W.; Henderson, W.A. Elevated circulating miR-150 and miR-342-3p in patients with irritable bowel syndrome. Exp. Mol. Pathol. 2014, 96, 422–425. [Google Scholar] [CrossRef] [PubMed]
  103. Zhou, Q.; Souba, W.W.; Croce, C.M.; Verne, G.N. MicroRNA-29a regulates intestinal membrane permeability in patients with irritable bowel syndrome. Gut 2009, 59, 775–784. [Google Scholar] [CrossRef] [PubMed]
  104. Hou, Q.; Huang, Y.; Zhu, S.; Li, P.; Chen, X.; Hou, Z.; Liu, F. MiR-144 increases intestinal permeability in IBS-D rats by targeting OCLN and ZO1. Cell. Physiol. Biochem. 2018, 44, 2256–2268. [Google Scholar] [CrossRef] [PubMed]
  105. Burney, R.; Hamilton, A.; Aghajanova, L.; Vo, K.; Nezhat, C.; Lessey, B.; Giudice, L. MicroRNA expression profiling of eutopic secretory endometrium in women with versus without endometriosis. Mol. Hum. Reprod. 2009, 15, 625–631. [Google Scholar] [CrossRef] [PubMed]
  106. Kästingschäfer, C.S.; Schäfer, S.D.; Kiesel, L.; Götte, M. miR-142-3p is a novel regulator of cell viability and proinflammatory signalling in endometrial stroma cells. Reprod. BioMedicine Online 2015, 30, 553–556. [Google Scholar] [CrossRef]
  107. Wright, K.R.; Mitchell, B.; Santanam, N. Redox regulation of microRNAs in endometriosis-associated pain. Redox Biol. 2017, 12, 956–966. [Google Scholar] [CrossRef] [PubMed]
  108. Maged, A.M.; Deeb, W.S.; El Amir, A.; Zaki, S.S.; El Sawah, H.; Al Mohamady, M.; Metwally, A.A.; Katta, M.A. Diagnostic accuracy of serum miR-122 and miR-199a in women with endometriosis. Int. J. Gynecol. Obstet. 2018, 141, 14–19. [Google Scholar] [CrossRef]
  109. Freire, V.S.; Burkhard, F.C.; Kessler, T.M.; Kuhn, A.; Draeger, A.; Monastyrskaya, K. MicroRNAs may mediate the down-regulation of neurokinin-1 receptor in chronic bladder pain syndrome. Am. J. Pathol. 2010, 176, 288–303. [Google Scholar] [CrossRef] [PubMed]
  110. Monastyrskaya, K.; Sánchez-Freire, V.; Gheinani, A.H.; Klumpp, D.J.; Babiychuk, E.B.; Draeger, A.; Burkhard, F.C. miR-199a-5p regulates urothelial permeability and may play a role in bladder pain syndrome. Am. J. Pathol. 2013, 182, 431–448. [Google Scholar] [CrossRef] [PubMed]
  111. Yan, Y.; Ma, Z.; Zhu, J.; Zeng, M.; Liu, H.; Dong, Z. miR-214 represses mitofusin-2 to promote renal tubular apoptosis in ischemic acute kidney injury. Am. J. Physiol.-Ren. Physiol. 2020, 318, F878–F887. [Google Scholar] [CrossRef]
  112. Wu, C.; Rakhshandehroo, T.; Wettersten, H.I.; Campos, A.; Von Schalscha, T.; Jain, S.; Yu, Z.; Tan, J.; Mose, E.; Childers, B.G. Pancreatic cancer cells upregulate LPAR4 in response to isolation stress to promote an ECM-enriched niche and support tumour initiation. Nat. Cell Biol. 2023, 25, 309–322. [Google Scholar] [CrossRef] [PubMed]
  113. Sengupta, J.N.; Pochiraju, S.; Kannampalli, P.; Bruckert, M.; Addya, S.; Yadav, P.; Miranda, A.; Shaker, R.; Banerjee, B. MicroRNA-mediated GABAAα-1 receptor subunit down-regulation in adult spinal cord following neonatal cystitis-induced chronic visceral pain in rats. Pain 2013, 154, 59–70. [Google Scholar] [CrossRef] [PubMed]
  114. Zhang, J.; Yu, J.; Kannampalli, P.; Nie, L.; Meng, H.; Medda, B.K.; Shaker, R.; Sengupta, J.N.; Banerjee, B. MicroRNA–mediated downregulation of potassium-chloride-cotransporter and vesicular γ-aminobutyric acid transporter expression in spinal cord contributes to neonatal cystitis–induced visceral pain in rats. Pain 2017, 158, 2461–2474. [Google Scholar] [CrossRef] [PubMed]
  115. Yam, M.F.; Loh, Y.C.; Tan, C.S.; Khadijah Adam, S.; Abdul Manan, N.; Basir, R. General pathways of pain sensation and the major neurotransmitters involved in pain regulation. Int. J. Mol. Sci. 2018, 19, 2164. [Google Scholar] [CrossRef] [PubMed]
  116. Pertovaara, A. It’s not cool to reduce the skin temperature and activate the TRPM8 ion channel after spinal injury. Scand. J. Pain 2013, 4, 31–32. [Google Scholar] [CrossRef] [PubMed]
  117. Haggard, P.; Iannetti, G.D.; Longo, M.R. Spatial sensory organization and body representation in pain perception. Current Biology 2013, 23, R164–R176. [Google Scholar] [CrossRef] [PubMed]
  118. Theodosis, D.T.; Poulain, D.A.; Oliet, S.H. Activity-dependent structural and functional plasticity of astrocyte-neuron interactions. Physiol. Rev. 2008, 88, 983–1008. [Google Scholar] [CrossRef] [PubMed]
  119. Craig, K.D.; Versloot, J.; Goubert, L.; Vervoort, T.; Crombez, G. Perceiving pain in others: Automatic and controlled mechanisms. J. Pain 2010, 11, 101–108. [Google Scholar] [CrossRef] [PubMed]
  120. Zhou, H.-Y.; Chen, S.-R.; Pan, H.-L. Targeting N-methyl-D-aspartate receptors for treatment of neuropathic pain. Expert Rev. Clin. Pharmacol. 2011, 4, 379–388. [Google Scholar] [CrossRef] [PubMed]
  121. Yan, X.-T.; Ji, L.-J.; Wang, Z.; Wu, X.; Wang, Q.; Sun, S.; Lu, J.-M.; Zhang, Y. MicroRNA-93 alleviates neuropathic pain through targeting signal transducer and activator of transcription 3. Int. Immunopharmacol. 2017, 46, 156–162. [Google Scholar] [CrossRef] [PubMed]
  122. Lo, W.-Y.; Peng, C.-T.; Wang, H.-J. MicroRNA-146a-5p mediates high glucose-induced endothelial inflammation via targeting interleukin-1 receptor-associated kinase 1 expression. Front. Physiol. 2017, 8, 551. [Google Scholar] [CrossRef] [PubMed]
  123. Elramah, S.; Landry, M.; Favereaux, A. MicroRNAs regulate neuronal plasticity and are involved in pain mechanisms. Front. Cell. Neurosci. 2014, 8, 31. [Google Scholar] [CrossRef] [PubMed]
  124. Niederberger, E.; Kynast, K.; Lötsch, J.; Geisslinger, G. MicroRNAs as new players in the pain game. Pain 2011, 152, 1455–1458. [Google Scholar] [CrossRef] [PubMed]
  125. Yi, L.-T.; Li, J.; Liu, B.-B.; Luo, L.; Liu, Q.; Geng, D. BDNF–ERK–CREB signalling mediates the role of miR-132 in the regulation of the effects of oleanolic acid in male mice. J. Psychiatry Neurosci. 2014, 39, 348–359. [Google Scholar] [CrossRef] [PubMed]
  126. Fortin, D.A.; Srivastava, T.; Soderling, T.R. Structural modulation of dendritic spines during synaptic plasticity. The Neuroscientist 2012, 18, 326–341. [Google Scholar] [CrossRef] [PubMed]
  127. Hu, Z.; Li, Z. miRNAs in synapse development and synaptic plasticity. Curr. Opin. Neurobiol. 2017, 45, 24–31. [Google Scholar] [CrossRef] [PubMed]
  128. Reza-Zaldivar, E.E.; Hernández-Sápiens, M.A.; Minjarez, B.; Gómez-Pinedo, U.; Sánchez-González, V.J.; Márquez-Aguirre, A.L.; Canales-Aguirre, A.A. Dendritic spine and synaptic plasticity in Alzheimer’s disease: A focus on MicroRNA. Front. Cell Dev. Biol. 2020, 8, 255. [Google Scholar] [CrossRef] [PubMed]
  129. López-González, M.J.; Landry, M.; Favereaux, A. MicroRNA and chronic pain: From mechanisms to therapeutic potential. Pharmacol. Ther. 2017, 180, 1–15. [Google Scholar] [CrossRef] [PubMed]
  130. Lutz, B.M.; Bekker, A.; Tao, Y.-X. Noncoding RNAs: New players in chronic pain. Anesthesiology 2014, 121, 409–417. [Google Scholar] [CrossRef] [PubMed]
  131. Bai, G.; Ambalavanar, R.; Wei, D.; Dessem, D. Downregulation of selective microRNAs in trigeminal ganglion neurons following inflammatory muscle pain. Mol. Pain 2007, 3, 1–4. [Google Scholar] [CrossRef] [PubMed]
  132. Recchiuti, A.; Krishnamoorthy, S.; Fredman, G.; Chiang, N.; Serhan, C.N. MicroRNAs in resolution of acute inflammation: Identification of novel resolvin D1-miRNA circuits. FASEB J. 2011, 25, 544. [Google Scholar] [CrossRef] [PubMed]
  133. Xu, Z.-Z.; Zhang, L.; Liu, T.; Park, J.Y.; Berta, T.; Yang, R.; Serhan, C.N.; Ji, R.-R. Resolvins RvE1 and RvD1 attenuate inflammatory pain via central and peripheral actions. Nat. Med. 2010, 16, 592–597. [Google Scholar] [CrossRef] [PubMed]
  134. Wesselmann, U.; Baranowski, A.P.; Börjesson, M.; Curran, N.C.; Czakanski, P.P.; Giamberardino, M.A.; Ness, T.J.; Robbins, M.T.; Traub, R.J. Emerging therapies and novel approaches to visceral pain. Drug Discov. Today Ther. Strateg. 2009, 6, 89–95. [Google Scholar] [CrossRef] [PubMed]
  135. Ji, R.-R.; Kohno, T.; Moore, K.A.; Woolf, C.J. Central sensitization and LTP: Do pain and memory share similar mechanisms? Trends Neurosci. 2003, 26, 696–705. [Google Scholar] [CrossRef]
  136. Bravo-Vázquez, L.A.; Medina-Ríos, I.; Márquez-Gallardo, L.D.; Reyes-Muñoz, J.; Serrano-Cano, F.I.; Pathak, S.; Paul, S. Functional implications and clinical potential of MicroRNAs in irritable bowel syndrome: A concise review. Dig. Dis. Sci. 2023, 68, 38–53. [Google Scholar] [CrossRef] [PubMed]
  137. Lin, Z.; Hu, H.; Liu, B.; Chen, Y.; Tao, Y.; Zhou, X.; Li, M. Biomaterial-assisted drug delivery for interstitial cystitis/bladder pain syndrome treatment. J. Mater. Chem. B 2021, 9, 23–34. [Google Scholar] [CrossRef] [PubMed]
  138. Fassbender, A.; O, D.; De Moor, B.; Waelkens, E.; Meuleman, C.; Tomassetti, C.; Peeraer, K.; D’Hooghe, T. Biomarkers of endometriosis. Fertil Steril. 2013, 99, 1135–1145. [Google Scholar] [CrossRef] [PubMed]
  139. Bjorkman, S.; Taylor, H.S. MicroRNAs in endometriosis: Biological function and emerging biomarker candidates. Biol. Reprod. 2019, 101, 1167–1178. [Google Scholar] [CrossRef] [PubMed]
  140. Barker, J.S.; Hines, R.M. Regulation of GABAA receptor subunit expression in substance use disorders. Int. J. Mol. Sci. 2020, 21, 4445. [Google Scholar] [CrossRef] [PubMed]
  141. Vasudevan, S. Posttranscriptional upregulation by microRNAs. Wiley Interdiscip. Rev. RNA 2012, 3, 311–330. [Google Scholar] [CrossRef] [PubMed]
  142. Kress, M.; Hüttenhofer, A.; Landry, M.; Kuner, R.; Favereaux, A.; Greenberg, D.S.; Bednarik, J.; Heppenstall, P.; Kronenberg, F.; Malcangio, M. microRNAs in nociceptive circuits as predictors of future clinical applications. Front. Mol. Neurosci. 2013, 6, 33. [Google Scholar] [CrossRef] [PubMed]
  143. Xiao, Y.; Liu, H.; Chen, L.; Wang, Y.; Yao, X.; Jiang, X. Association of microRNAs genes polymorphisms with arthritis: A systematic review and meta-analysis. Biosci. Rep. 2019, 39, BSR20190298. [Google Scholar] [CrossRef]
  144. Jiangpan, P.; Qingsheng, M.; Zhiwen, Y.; Tao, Z. Emerging role of microRNA in neuropathic pain. Curr. Drug Metab. 2016, 17, 336–344. [Google Scholar] [CrossRef] [PubMed]
  145. Bali, K.K.; Kuner, R. Noncoding RNAs: Key molecules in understanding and treating pain. Trends Mol. Med. 2014, 20, 437–448. [Google Scholar] [CrossRef]
  146. Tan, P.-H.; Pao, Y.-Y.; Cheng, J.-K.; Hung, K.-C.; Liu, C.-C. MicroRNA-based therapy in pain medicine: Current progress and future prospects. Acta Anaesthesiol. Taiwanica 2013, 51, 171–176. [Google Scholar] [CrossRef]
  147. Garzon, R.; Marcucci, G.; Croce, C.M. Targeting microRNAs in cancer: Rationale, strategies and challenges. Nat. Rev. Drug Discov. 2010, 9, 775–789. [Google Scholar] [CrossRef] [PubMed]
  148. Dreyer, J.-L. Lentiviral vector-mediated gene transfer and RNA silencing technology in neuronal dysfunctions. In Lentivirus Gene Engineering Protocols, 2nd ed.; Springer: Berlin/Heidelberg, Germany, 2010; pp. 3–35. [Google Scholar]
  149. Sun, B.; Gan, L. Manipulation of gene expression in the central nervous system with lentiviral vectors. In Alzheimer’s Disease and Frontotemporal Dementia: Methods and Protocols; Springer: Berlin/Heidelberg, Germany, 2011; pp. 155–168. [Google Scholar]
  150. Bäumer, N.; Berdel, W.E.; Bäumer, S. Immunoprotein-mediated siRNA delivery. Mol. Pharm. 2017, 14, 1339–1351. [Google Scholar] [CrossRef] [PubMed]
  151. Campos Pereira, T.; Lopes-Cendes, I. Emerging RNA-based drugs: siRNAs, microRNAs and derivates. Cent. Nerv. Syst. Agents Med. Chem. (Former. Curr. Med. Chem.-Cent. Nerv. Syst. Agents) 2012, 12, 217–232. [Google Scholar] [CrossRef] [PubMed]
  152. Ramanathan, S.; Ajit, S.K. MicroRNA-based biomarkers in pain. Adv. Pharmacol. 2016, 75, 35–62. [Google Scholar] [PubMed]
  153. Al-Rawaf, H.A.; Alghadir, A.H.; Gabr, S.A. MicroRNAs as biomarkers of pain intensity in patients with chronic fatigue syndrome. Pain Pract. 2019, 19, 848–860. [Google Scholar] [CrossRef] [PubMed]
  154. Sakai, A.; Suzuki, H. microRNA and Pain. In microRNA: Medical Evidence: From Molecular Biology to Clinical Practice; Springer: Berlin/Heidelberg, Germany, 2015; pp. 17–39. [Google Scholar]
  155. Kusuda, R.; Cadetti, F.; Ravanelli, M.I.; Sousa, T.A.; Zanon, S.; De Lucca, F.L.; Lucas, G. Differential expression of microRNAs in mouse pain models. Mol. Pain 2011, 7, 1–8. [Google Scholar] [CrossRef] [PubMed]
  156. Li, X.; Kroin, J.S.; Kc, R.; Gibson, G.; Chen, D.; Corbett, G.T.; Pahan, K.; Fayyaz, S.; Kim, J.S.; Van Wijnen, A.J. Altered spinal microRNA-146a and the microRNA-183 cluster contribute to osteoarthritic pain in knee joints. J. Bone Miner. Res. 2013, 28, 2512–2522. [Google Scholar] [CrossRef] [PubMed]
  157. Kynast, K.L.; Russe, O.Q.; Geisslinger, G.; Niederberger, E. Novel findings in pain processing pathways: Implications for miRNAs as future therapeutic targets. Expert Rev. Neurother. 2013, 13, 515–525. [Google Scholar] [CrossRef] [PubMed]
  158. Ciszek, B.P.; Khan, A.A.; Dang, H.; Slade, G.D.; Smith, S.; Bair, E.; Maixner, W.; Zolnoun, D.; Nackley, A.G. MicroRNA expression profiles differentiate chronic pain condition subtypes. Transl. Res. 2015, 166, 706–720.e711. [Google Scholar] [CrossRef] [PubMed]
  159. Maharshak, N.; Shenhar-Tsarfaty, S.; Aroyo, N.; Orpaz, N.; Guberman, I.; Canaani, J.; Halpern, Z.; Dotan, I.; Berliner, S.; Soreq, H. MicroRNA-132 modulates cholinergic signaling and inflammation in human inflammatory bowel disease. Inflamm. Bowel Dis. 2013, 19, 1346–1353. [Google Scholar] [CrossRef] [PubMed]
  160. Vaknine, S.; Soreq, H. Central and peripheral anti-inflammatory effects of acetylcholinesterase inhibitors. Neuropharmacology 2020, 168, 108020. [Google Scholar] [CrossRef] [PubMed]
  161. Yu, B.; Zhou, S.; Yi, S.; Gu, X. The regulatory roles of non-coding RNAs in nerve injury and regeneration. Prog. Neurobiol. 2015, 134, 122–139. [Google Scholar] [CrossRef] [PubMed]
  162. Burrows, K.; Figueroa-Hall, L.K.; Alarbi, A.M.; Stewart, J.L.; Kuplicki, R.; Tan, C.; Hannafon, B.N.; Ramesh, R.; Savitz, J.; Khalsa, S. Association between inflammation, reward processing, and ibuprofen-induced increases of miR-23b in astrocyte-enriched extracellular vesicles: A randomized, placebo-controlled, double-blind, exploratory trial in healthy individuals. Brain Behav. Immun.-Health 2023, 27, 100582. [Google Scholar] [CrossRef] [PubMed]
  163. Zhao, J.; He, Z.; Wang, J. MicroRNA-124: A key player in microglia-mediated inflammation in neurological diseases. Front. Cell. Neurosci. 2021, 15, 771898. [Google Scholar] [CrossRef] [PubMed]
  164. Zingale, V.D.; Gugliandolo, A.; Mazzon, E. MiR-155: An important regulator of Neuroinflammation. Int. J. Mol. Sci. 2021, 23, 90. [Google Scholar] [CrossRef] [PubMed]
  165. Liao, Z.; Zheng, R.; Shao, G. Mechanisms and application strategies of miRNA-146a regulating inflammation and fibrosis at molecular and cellular levels. Int. J. Mol. Med. 2023, 51, 1–16. [Google Scholar] [CrossRef] [PubMed]
  166. Heilig, M.; Barbier, E.; Johnstone, A.; Tapocik, J.; Meinhardt, M.; Pfarr, S.; Wahlestedt, C.; Sommer, W. Reprogramming of mPFC transcriptome and function in alcohol dependence. Genes Brain Behav. 2017, 16, 86–100. [Google Scholar] [CrossRef] [PubMed]
  167. Cacabelos, R.; Torrellas, C. Epigenetics of aging and Alzheimer’s disease: Implications for pharmacogenomics and drug response. Int. J. Mol. Sci. 2015, 16, 30483–30543. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Role of miRs in altered gene expression and pain receptors [49]. Diagram showing the targets of miRs that are dysregulated in neuropathic pain and how they are implicated in intracellular signaling. The miRs (miR144, miR206, miR93, miR98, miR544, and miR206) are downregulated during neuropathic pain, according to red downward arrows; in contrast, upward green arrows indicate the elevation of miRs (miR15, miR16, miR155, miR221, and miR7a) during neuropathic pain.
Figure 1. Role of miRs in altered gene expression and pain receptors [49]. Diagram showing the targets of miRs that are dysregulated in neuropathic pain and how they are implicated in intracellular signaling. The miRs (miR144, miR206, miR93, miR98, miR544, and miR206) are downregulated during neuropathic pain, according to red downward arrows; in contrast, upward green arrows indicate the elevation of miRs (miR15, miR16, miR155, miR221, and miR7a) during neuropathic pain.
Genes 15 01003 g001
Figure 2. miR-199, TRPV1, visceral hypersensitivity, and visceral pain may have a mechanistic link. Several potentially special benefits over traditional retroviral gene delivery methods are provided by lentiviruses, which deliver miRs in vivo. The primary functions of these are to infect non-dividing cells, including dorsal root ganglion neuron cells, and to produce stable, long-lasting gene/miRNA expression. This work delivered and overexpressed miR-199a to suppress the expression of TRPV1 using lentiviruses. Pictures depict the associated molecular pathways. Lower left panel: Stressors in the environment cause damage to the stomach, which starts a chain reaction that makes the body hypersensitive to certain stimulants. A reduction in miR-199 and an increase in TRPV1 expression could be indicative of this hypersensitive condition [99].
Figure 2. miR-199, TRPV1, visceral hypersensitivity, and visceral pain may have a mechanistic link. Several potentially special benefits over traditional retroviral gene delivery methods are provided by lentiviruses, which deliver miRs in vivo. The primary functions of these are to infect non-dividing cells, including dorsal root ganglion neuron cells, and to produce stable, long-lasting gene/miRNA expression. This work delivered and overexpressed miR-199a to suppress the expression of TRPV1 using lentiviruses. Pictures depict the associated molecular pathways. Lower left panel: Stressors in the environment cause damage to the stomach, which starts a chain reaction that makes the body hypersensitive to certain stimulants. A reduction in miR-199 and an increase in TRPV1 expression could be indicative of this hypersensitive condition [99].
Genes 15 01003 g002
Figure 3. Polymorphism effect of miRs. The figure shows the formation of miRs and their polymorphism effects, like translation repression and DNA degradation, by targeting mRNA.
Figure 3. Polymorphism effect of miRs. The figure shows the formation of miRs and their polymorphism effects, like translation repression and DNA degradation, by targeting mRNA.
Genes 15 01003 g003
Table 1. Role of miRs in pain receptors [98].
Table 1. Role of miRs in pain receptors [98].
DiseasesmiRsTissuesTargetsRef
IBSmiR-199 ↓Human colon; Rat colon/DRGTRPV1 ↑[99]
miR-24 ↑Human/mouse intestinal mucosaSERT ↓[100]
miR-17-5p ↑Lumbar spinal cordSTAT3 ↓; gp130 ↑[101]
miR-150 ↑
miR-342-3p ↑
Human whole blood-[102]
miR-29a ↑Human small bowel and colon; human blood macrovesiclesGlutamate ammonia
ligase ↓
[103]
miR-144 ↑Rat distal colonic epithelial cellsOccludin ↓; ZO1 ↓[104]
EndometriosismiR-9 ↓;
miR-34 ↓
Human endometrial tissues-[105]
miR-142-3p ↑Endometrial stroma cellsSteroid sulfatase ↓;
gp130 ↓
[106]
miR-29 ↑;
miR-181 ↑;
let-7 ↑
ox-LDL-treated human endometrial cell linesNGF ↑; IL-6 ↑;
PTGES3 ↑
[107]
miR-122 ↑;
miR-199a ↑
Serum; peritoneal fluid-[108]
BPS/ICmiR-449b ↑
miR-500 ↑
Bladder smooth muscle cellsNK1 receptor↓[109]
miR-199a-5p ↑Bladder smooth muscle;
Mature bladder urothelium;
Primary urothelial culture
LIN7C ↓;
ARHGAP12 ↓;
PALS1 ↓;
RND1↓; PVRL1 ↓
[110]
miR-214 ↓Postmenopausal women’s bladder tissue;
Ovariectomized rats’ APMSCs
Mfn2 ↑[111]
miR-139-5p ↓Postmenopausal women’s bladder tissueLPAR4 ↑[112]
miR-181a ↑Rat spinal cordGABAA[113]
miR-92b-3p ↑Rat spinal cordKCC2 ↓; VGAT ↓[114]
Table 2. miRNAs and gene association.
Table 2. miRNAs and gene association.
miRsTargetsRoleRef
miR-1Sodium voltage-gated channel alpha subunit 1 (SCN1A)Involved in controlling neuronal excitability, and neuropathic pain has been linked to its dysfunction[49]
miR-21Programmed cell death 4 (PDCD4), Sprouty homolog 2 (SPRY2), and othersImplicated in neuroinflammation and it has been found to be upregulated in models of neuropathic pain[161]
miR-23bProstaglandin-endoperoxide synthase 2 (PTGS2/COX-2)Involved in controlling inflammatory pathways; a malfunction in this regard could be the cause of inflammatory pain[162]
miR-124Signal transducer and activator of transcription 3 (STAT3)Linked to neuroinflammation and microglial activation; dysregulation may be a factor in neuropathic pain[163]
miR-155Suppressor of cytokine signaling 1 (SOCS1), SH2-containing inositol phosphatase 1 (SHIP1), and othersInvolved in immune response and neuroinflammatory modulation; elevated in chronic pain models[164]
miR-146aInterleukin-1 receptor-associated kinase 1 (IRAK1) and tumor necrosis factor receptor-associated factor 6 (TRAF6)Involved in the regulation of immune responses and inflammation; dysregulation may contribute to chronic pain conditions[165]
miR-29Collagens, involved in extracellular matrix regulationLinked to the control of extracellular matrix components and fibromyalgia; dysregulation may be a factor in the abnormalities of connective tissue in chronic pain[124]
miR-30aSerine/threonine-protein kinase WNK1Connected to WNK1 expression variation, which may affect sensitivity to pain[166]
miR-128Voltage-gated sodium channel alpha subunit 2 (SCN2A)Involved in controlling the excitability of neurons; dysregulation could lead to neuropathic pain[167]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Hassan, M.; Shahzadi, S.; Yasir, M.; Chun, W.; Kloczkowski, A. Therapeutic Implication of miRNAs as an Active Regulatory Player in the Management of Pain: A Review. Genes 2024, 15, 1003. https://doi.org/10.3390/genes15081003

AMA Style

Hassan M, Shahzadi S, Yasir M, Chun W, Kloczkowski A. Therapeutic Implication of miRNAs as an Active Regulatory Player in the Management of Pain: A Review. Genes. 2024; 15(8):1003. https://doi.org/10.3390/genes15081003

Chicago/Turabian Style

Hassan, Mubashir, Saba Shahzadi, Muhammad Yasir, Wanjoo Chun, and Andrzej Kloczkowski. 2024. "Therapeutic Implication of miRNAs as an Active Regulatory Player in the Management of Pain: A Review" Genes 15, no. 8: 1003. https://doi.org/10.3390/genes15081003

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