Hostname: page-component-76fb5796d-skm99 Total loading time: 0 Render date: 2024-04-25T16:18:34.671Z Has data issue: false hasContentIssue false

The Known Biology of Neuropathic Pain and Its Relevance to Pain Management

Published online by Cambridge University Press:  17 February 2023

Peter A. Smith*
Affiliation:
Neuroscience and Mental Health Institute and Department of Pharmacology, University of Alberta, Edmonton, Canada
*
Corresponding author: Peter A. Smith, Ph.D., Neuroscience and Mental Health Institute and Department of Pharmacology, Faculty of Medicine and Dentistry, University of Alberta, 9-70 Medical Sciences Building, Edmonton, AB, Canada, T6G 2H7. Email: pas3@ualberta.ca
Rights & Permissions [Opens in a new window]

Abstract:

Patients with neuropathic pain are heterogeneous in pathophysiology, etiology, and clinical presentation. Signs and symptoms are determined by the nature of the injury and factors such as genetics, sex, prior injury, age, culture, and environment. Basic science has provided general information about pain etiology by studying the consequences of peripheral injury in rodent models. This is associated with the release of inflammatory cytokines, chemokines, and growth factors that sensitize sensory nerve endings, alter gene expression, promote post-translational modification of proteins, and alter ion channel function. This leads to spontaneous activity in primary afferent neurons that is crucial for the onset and persistence of pain and the release of secondary mediators such as colony-stimulating factor 1 from primary afferent terminals. These promote the release of tertiary mediators such as brain-derived neurotrophic factor and interleukin-1β from microglia and astrocytes. Tertiary mediators facilitate the transmission of nociceptive information at the spinal, thalamic, and cortical levels. For the most part, these findings have failed to identify new therapeutic approaches. More recent basic science has better mirrored the clinical situation by addressing the pathophysiology associated with specific types of injury, refinement of methodology, and attention to various contributory factors such as sex. Improved quantification of sensory profiles in each patient and their distribution into defined clusters may improve translation between basic science and clinical practice. If such quantification can be traced back to cellular and molecular aspects of pathophysiology, this may lead to personalized medicine approaches that dictate a rational therapeutic approach for each individual.

Résumé :

RÉSUMÉ :

Nos connaissances actuelles en biologie en ce qui concerne la douleur neuropathique et leur pertinence pour la prise en charge de la douleur.

Les patients souffrant de douleurs neuropathiques sont hétérogènes en termes de pathophysiologie, d’étiologie et de présentation clinique. Leurs signes cliniques et leurs symptômes sont déterminés par la nature de leur lésion ainsi que par des facteurs tels que la génétique, le sexe, une lésion antérieure, l’âge, la culture et l’environnement. À l’aide de modèles appliqués à des rongeurs, nos connaissances scientifiques fondamentales ont fourni des éléments généraux d’information sur l’étiologie de la douleur en étudiant les conséquences de lésions périphériques. Un tel phénomène est associé à la libération de cytokines pro-inflammatoires, de chimiokines et de facteurs de croissance qui sensibilisent les terminaisons nerveuses sensorielles, modifient l’expression génétique, favorisent la modification post-traductionnelle des protéines et altèrent la fonction des canaux ioniques. Cela conduit en retour à une activité spontanée dans les neurones afférents primaires, laquelle est cruciale dans l’apparition et la persistance de la douleur et la libération de médiateurs secondaires, par exemple le récepteur de « facteur de stimulation des colonies 1 » à partir des terminaisons afférentes primaires. Ces dernières favorisent par ailleurs la libération de médiateurs tertiaires tels que le facteur neurotrophique dérivé du cerveau (FNDC) et l’interleukine-1β par la microglie et les astrocytes. Les médiateurs tertiaires facilitent aussi la transmission des informations nociceptives aux niveaux spinal, thalamique et cortical. Dans l’ensemble, ces découvertes n’ont pas permis d’identifier de nouvelles approches thérapeutiques. Cela dit, les avancées plus récentes de la science fondamentale reflètent mieux la situation clinique des patients en abordant la ou les pathophysiologies associées à des types spécifiques de lésions, en affinant la méthodologie employée et en prêtant attention à divers facteurs contributifs, par exemple le sexe. Une meilleure quantification du profil sensoriel de chaque patient et leur répartition en groupes définis peuvent ainsi améliorer le transfert entre les connaissances fondamentales de la science et la pratique clinique. Si cette quantification parvient à remonter jusqu’aux aspects cellulaires et moléculaires de la physiopathologie, cela pourrait conduire à des approches médicales personnalisées qui dictent une approche thérapeutique rationnelle pour chaque individu.

Type
Review Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2023. Published by Cambridge University Press on behalf of Canadian Neurological Sciences Federation

Introduction

The signs and symptoms of neuropathic pain in each individual patient are strongly dependent on variables such as sex, age, ethnicity, inherited genetic predisposition, intestinal microbiome, prior neonatal injury, personality variables, and cultural and environmental factors. Reference Mogil1Reference Baron, Maier and Attal5 This heterogeneity of presentation also reflects the association of neuropathic pain with a diverse set of maladies. These include peripheral nerve trauma, brain or spinal cord injury, fibromyalgia, multiple sclerosis, spinal, cortical or brain stem cord stroke, post herpetic and trigeminal neuralgia, migraine, osteoarthritis, rheumatoid arthritis, autoimmune disease, complex regional pain syndromes I and II, viral infections such as HIV and COVID-19 and neuropathies associated with diabetes, chemotherapy, and cancer itself. Reference Finnerup, Kuner and Jensen6

Signs and symptoms include hyperalgesia, mechanical, or cold-induced allodynia, bouts of spontaneous “electric shock like” pain and sometimes the persistent burning pain of causalgia Reference Finnerup, Kuner and Jensen6 Some patients experience sensory disturbances. These may involve paresthesias, described as a crawling sensation, pricking or tingling Reference Bannister, Sachau, Baron and Dickenson7 or anesthesia dolorosa where the area of injury is painful yet insensitive to touch. Reference Wall, Devor and Inbal8 Neuropathic pain is frequently intractable, relatively insensitive to the action of opioids Reference Alles and Smith9,Reference Finnerup, Attal and Haroutounian10 and may present with co-morbidities such as anxiety, irritability, sleep disorders, depression, and/or sensory abnormalities. Reference Bannister, Sachau, Baron and Dickenson7,Reference Gormsen, Rosenberg, Bach and Jensen11 Despite intensive efforts to find new drugs and targets over the past 30 years, the urgent need to find new treatments persists. Reference Finnerup, Kuner and Jensen6,Reference Alles and Smith9,Reference Finnerup, Attal and Haroutounian10,Reference Price, Basbaum and Bresnahan12 Most of the current understanding is derived from peripheral nerve injury models in rodents. In most cases, the spared nerve injury (SNI) or chronic constriction injury (CCI) models are used. Reference Decosterd and Woolf13

This review will overview the current understanding of pain induced in animal models by peripheral nerve injury. In view of the recognized knowledge gap between these basic science results and the various signs and symptoms and/or pain phenotypes seen in patients, Reference Price, Basbaum and Bresnahan12 a brief outline of clinical and basic science strategies that seek to bridge this gap will be presented.

Nerve Injury, Wallarian Degeneration, and Primary Mediators

Following injury, Wallerian degeneration of severed axons is associated with neutrophil, macrophage, and T-lymphocyte invasion as well as activation of Schwann cells, fibroblasts, mast cells, keratinocytes, and epithelial cells. Reference Scholz and Woolf14Reference Marais, Light, Mason, Paterson, Olson and Marshall16 Once activated, these immunocompetent cells generate and release pro-inflammatory primary mediators. These include tumor necrosis factor (TNF-α), Reference Scholz and Woolf14,Reference Leung and Cahill17 interleukins 1β,15,17 and 18 (IL-1β, IL-15, IL-17, and IL-18), Reference Scholz and Woolf14,Reference Boakye, Tang and Smith18Reference Kleinschnitz, Hofstetter, Meuth, Braeuninger, Sommer and Stoll21 nerve growth factor, Reference Scholz and Woolf14,Reference Pezet and McMahon22 monocyte chemoattractant protein 1 (MCP-1/CCL-2), Reference White and Wilson23 chemokine (C-X-C motif) ligands 1 (CXCL-1) Reference Scholz and Woolf14,Reference Silva, Lopes, Guimaraes and Cunha24 and 12 (CXCL-12), Reference Yu, Huang, Di, Qu and Fan25 histamine, prostaglandins, serotonin, and substance P Reference Scholz and Woolf14,Reference Kaur, Singh and Jaggi26,Reference ObaraI Telezhkin, Alrashdi and Chazot27 as well as the secreted glycoproteins Wnt3a (wingless-type mammary tumor virus integration site family member 3a) and Wnt5a. Reference van Vliet, Lee and van der Poel28

Structural Remodeling of Injured Peripheral Nerves

Following SNI of rodent peripheral nerves, degeneration of the axons of low threshold non-nociceptive afferents can lead to loss of sensation. Peripheral nociceptors then sprout into territories that were previously occupied by low threshold afferents. These nociceptors are transformed to exhibit a low activation threshold so that mild tactile stimulation now produces mechanical allodynia. Reference Gangadharan, Zheng and Tabener29

In many cases, injury also provokes the sprouting of perivascular sympathetic fibers so that they interact and excite sensory nerve terminals and DRG cell bodies. Reference McLachlan, Janig and Michalis30,Reference Yen, Bennett and Ribeiro-da-Silva31 These processes are especially relevant to the etiology of complex regional pain syndrome II. Reference Drummond, Drummond and Dawson32

Injury-Induced Peripheral Sensitization, the Importance of Spontaneous Activity, and the Generation of Secondary Mediators

Immune cell-derived primary mediators sensitize peripheral nerve endings, axons, and cell bodies of primary afferents. Reference Scholz and Woolf14 Mediators also promote plasma extravasation and increase the permeability of the blood–brain barrier Reference Xanthos, Pungel, Wunderbaldinger and Sandkuhler33 and the blood–nerve barrier in the periphery. Reference Lim, Shi and Martin34 This and the chemoattractant profiles of various mediators facilitate the recruitment of immunocompetent leucocytes and lymphocytes to the site of injury. Reference Moalem and Tracey15,Reference Gomez-Nicola, Valle-Argos, Suardiaz, Taylor and Nieto-Sampedro20 These myeloid and lymphoid cells themselves release a host of cytokines and chemokines thereby instigating a positive feedback process in the initiation and maintenance of neuroinflammation and pain. Neuroinflammation is defined as activation of the brain’s innate immune system in response to an inflammatory challenge. Reference DiSabato, Quan and Godbout35,Reference Milatovic, Zaja-Milatovic, Breyer, Aschner, Montine and Gupta36

Satellite glial cells and resident macrophages in DRG Reference Noh, Mikler, Joy and Smith37Reference Xie, Strong and Zhang39 represent yet another source of inflammatory mediators. The actions of primary mediators such as IL-1β and TNF-α on DRG neurons culminate in marked changes in genes coding for neuropeptides, cytokines, chemokines, receptors, ion channels, signal transduction molecules, and synaptic vesicle proteins. Reference Zhang and Xiao40,Reference Biber and Boddeke41 Some of these gene products also function as secondary mediators that are released and effect the transfer of information between damaged peripheral nerves and various cell types in the spinal dorsal horn. Reference Boakye, Tang and Smith18

Primary mediators also control the expression of long non-coding RNA’s Reference Baskozos42 and microRNA’s in DRG. The latter are also upregulated by nerve injury Reference Finnerup, Kuner and Jensen6 and post-transcriptionally regulate the protein expression of hundreds of genes in a sequence-specific manner. Reference Liu, Xu and Wang43 Transfer of microRNAs between cell types may be brought about by the release and uptake of exosomes. Reference Zhang, Ye and Zhao44

Importantly, altered function of ion channels as a result of the action of primary mediators leads to increased excitability of primary afferent neurons Reference Noh, Stemkowski and Smith45Reference Chen, Pang and Shen49 and the generation of stimulus-independent spontaneous activity. This incessant spontaneous activity in primary afferents is absolutely crucial for the onset and persistence of pain. Reference Pitcher and Henry50Reference Yatziv and Devor53 This is illustrated by the effectiveness of topically applied lidocaine in the clinic. Reference Dworkin, O’Connor and Backonja54 Altered ion channel function and peripheral hyperexcitability may even be involved in spinal cord injury Reference Ritter, Zemel, Hala, O’Leary, Lepore and Covarrubais55 and central post-stroke pain. Reference Haroutounian, Ford and Frey56 Although Nav1.7, Kv7.2, Cav2.2, Cav3.2, and HCN2 channels have emerged as potential therapeutic targets for drug development, with the notable exception of gabapentinoid action on voltage-gated Ca2+ channels, Reference Alles and Smith9 pharmacological manipulation of these channels has failed to identify new therapeutic approaches. Reference Alles and Smith57

The observation that peripherally generated pain is often not suppressed by rhizotomy Reference Eschenfelder, Habler and Jannig58 seems at odds with the idea that stimulus-independent spontaneous activity is required for pain maintenance. It is possible, however, that pain seen after rhizotomy is related to deafferentation. This deafferentation pain may replace that which previously resulted from ectopic primary afferent activity. Reference Eschenfelder, Habler and Jannig58

As would be expected, the population of ion channels affected by primary mediators is similar to that affected by peripheral nerve injury Reference Noh, Stemkowski and Smith45,Reference Stemkowski, Noh, Chen and Smith47,Reference Stemkowski and Smith59 and in animal models, blockade of the actions of primary mediators abrogates signs of injury-induced pain. Reference Scholz and Woolf14,Reference Moalem and Tracey15,Reference Boakye, Tang and Smith18,Reference Wolf, Gabay, Tal, Yirmiya and Shavit60Reference Grace, Hutchinson, Maier and Watkins63 In general however, attempts to block the action of inflammatory mediators to limit neuropathic pain in the clinic have met with limited success. Reference Yekkirala, Roberson, Bean and Woolf64

Bidirectional Signalling between the Nervous and Immune Systems and “Neurogenic Neuroinflammation”

The relationship between immune cells and neurons is bidirectional. In addition to the well-documented actions of immune mediators on neurons, Reference Boakye, Tang and Smith18,Reference Noh, Stemkowski and Smith45Reference Binshtok, Wang and Zimmermann48,Reference Gustafson-Vickers, Lu, Lai, Todd, Ballanyi and Smith65Reference Vikman, Siddall and Duggan67 neuronal activity has a direct effect on immune cells. Reference Talbot, Foster and Woolf68Reference McMahon, La Russa and Bennett72 This “neurogenic neuroinflammation” Reference Xanthos and Sandkuhler73 is brought about by the release of neuropeptides and glutamate from primary afferents and their interaction with their cognate receptors on immune cells, astrocytes, and microglia. Reference McMahon, La Russa and Bennett72,Reference Shi, Wang, Li, Peymen, Kingerly and Clark74

Actions of Secondary Mediators and Transfer of Information from the Periphery to the Spinal Cord

Most secondary mediators are released from primary afferent terminals. Substances such as colony-stimulating factor 1 (CSF-1), the chemokines CCL-21, CXCL-12, and Wnt3a and Wnt5a Reference Finnerup, Kuner and Jensen6,Reference Boakye, Tang and Smith18,Reference van Vliet, Lee and van der Poel28,Reference Okubo, Yamanaka and Kobayashi75Reference Dong, Xu, Xia and Zhang78 activate their cognate receptors on spinal microglia and/or astrocytes and alter their properties. Activated glia thereby detects and mount an enduring response to peripheral nerve injury. Spinal microglia are affected in male rodents Reference Malcangio77 whereas invading macrophages and adaptive immune cells such as T-lymphocytes are involved in females. Reference Halievski, Ghazisaeidi and Salter79Reference Sorge and Mapplebeck81 CCL-21 and CXCL-12 signal to activate astrocytes. Reference Dong, Xu, Xia and Zhang78,Reference van Weering, de Jong, de Haas, Biber and Boddeke82 The inflammatory mediator, IFN-γ is increased in spinal cord following peripheral nerve injury Reference Costigan, Moss and Latremoliere83 and this may originate from invading T-lymphocytes.

Generation and Release of Tertiary Mediators in the Dorsal Horn and Central Sensitizaton

Glial activation and proliferation leads to the generation and release of tertiary mediators such as BDNF, IL-1β, TNF-α, and IFN-γ. Reference Boakye, Tang and Smith18,Reference Biggs, Lu, Stebbing, Balasubramanyan and Smith84,Reference Smith85

BDNF is released from microglia in response to the secondary mediators CSF-1 Reference Boakye, Tang and Smith18,Reference Guan, Kuhn and Wang76,Reference Yu, Basbaum and Guan86,Reference Boakye, Rancic and Whitlock87 and/or Wnt3a. Reference Zhang W.Shi and Peng88 BDNF release requires activation of P2X4 receptors by ATP. Reference Alles and Smith9,Reference Trang, Beggs, Wan and Salter89 As a mediator of the effect of nerve injury, Reference Coull, Boudreau and Bachand90Reference Chen, Balasubramanyan, Lai, Todd and Smith92 BDNF facilitates excitation Reference Biggs, Lu, Stebbing, Balasubramanyan and Smith84,Reference Lu, Ballanyi K.Colmers and Smith93Reference Hildebrand, Jian and Dedek95 and attenuates inhibition in the superficial dorsal horn. Reference Alles and Smith9,Reference Coull, Beggs and Boudreau96 These changes, which lead to central sensitization, spontaneous activity, and the misprocessing of sensory information, Reference Peirs and Seal97Reference Prescott, Ma and De Koninck100 involve at least four cellular mechanisms.

Microglial-derived BDNF increases excitatory drive to excitatory dorsal horn neurons and inhibits that to inhibitory neurons by both presynaptic and postsynaptic mechanisms. Reference Boakye, Rancic and Whitlock87,Reference Lu, Ballanyi K.Colmers and Smith93,Reference Lu, Biggs and Stebbing94 This altered synaptic activity is capable of increasing spontaneous action potential discharge in excitatory neurons while reducing it in inhibitory neurons. Reference Lu, Ballanyi K.Colmers and Smith93

BDNF also enhances excitatory responses to N-methyl-d aspartate (NMDA) in rat spinal cord in vitro. Reference Kerr, Bradbury and Bennett101 This may involve potentiation of the function of presynaptic NMDA receptors on primary afferent terminals Reference Chen, Walwyn and Ennes102 with a resultant increase in excitatory glutamatergic transmission. This may contribute to the effectiveness of the NMDA blocker, ketamine in some patients. Reference Dworkin, O’Connor and Backonja54

Peripheral nerve injury reduces expression of the potassium-chloride exporter (KCC2) selectively in nociceptive dorsal horn neurons. Reference Coull, Boudreau and Bachand90,Reference Ferrini, Perez-Sanchez and Ferland103 The resulting accumulation of intracellular Cl normally causes outward, inhibitory GABAergic synaptic currents mediated by Cl influx to become inward excitatory currents mediated by Cl efflux. Reference Coull, Boudreau and Bachand90 In male rats, this downregulation of KCC2 is mediated by BDNF. Reference Ferrini and De Koninck104 Since the loss of GABAergic inhibition enables non-noxious Aβ fiber-mediated excitatory transmission to access the superficial spinal dorsal horn, this process contributes to the establishment of allodynia. Reference Baba, Ji and Kohno99

Long-term potentiation (LTP) of synaptic transmission, sometimes known as “wind-up”, contributes to central sensitization in the dorsal horn. Reference Sandkuhler, Benrath, Brechtel, Ruscheweyh and Heinke105,Reference Sandkuhler106 LTP of C-fibre responses is augmented by BDNF Reference Li and Cai107 and LTP induced by nerve stimulation is occluded by BDNF pretreatment. Reference Ding, Cia and li108 The importance of these effects was recently underlined by the observation that spinal LTP as well as microglial activation and upregulation of BDNF are inhibited by antibodies to the secondary mediator CSF-1. This strongly implicates the CSF-1-microglia-BDNF axis Reference Boakye, Tang and Smith18 in the generation of spinal LTP. Reference Zhou, Peng and Xu109

As already mentioned, in females, changes in sensory processing in the dorsal horn involve the invasion of macrophages and T-lymphocytes. Reference Mapplebeck, Beggs and Salter80,Reference Sorge and Mapplebeck81 Yet as in males, this leads to attenuation of inhibition following the collapse of the Cl gradient. Reference Mapplebeck, Lorenzo and Lee110 In females, collapse of the Cl gradient is also brought about by the neuropeptide, CGRP Reference Paige, Plasencia-Fernandez and Kume111 which is released from primary afferent terminals. Reference Gardell, Vanderah and Gardell112

IL-1β from microglia stimulates astrocytic production of both TNF-α and IL-1β itself Reference Gajtko, Bakk, Hegedus, Ducza and Hollo113 thereby amplifying the initial IL-1β signal. Spinal actions of IL-1β involve increases in excitatory synaptic transmission. Reference Gustafson-Vickers, Lu, Lai, Todd, Ballanyi and Smith65,Reference Kawasaki, Zhang, Cheng and Ji66 This may involve a reduction in the ability of astrocytes to take up glutamate as a result of internalization of the astrocytic glutamate transporter (EAAT2). Reference Yan, Maixner and etal114

TNF-α also augments excitatory transmission in the dorsal horn Reference Boakye, Tang and Smith18,Reference Kawasaki, Zhang, Cheng and Ji66 as well as LTP by an action on glial cells. Reference Gruber-Schoffnegger, Drdla-Schutting, Honigsperger, Wunderbaldinger, Gassner and Sandkuhler115 Blockade of TNF-1 receptors attenuates neuropathic pain in male rodents but not in females. Reference del Rivero, Fischer, Yang, Swanson and Bethea116 Although anti-TNF antibodies and anti-TNF drugs such as thalidomide are available, none seem particularly useful in pain management. Reference Goncalves Dos, Delay, Yaksh and Corr117

IFN-γ from invading T-lymphocytes induces both tactile allodynia and altered microglia function. Genetic ablation of the interferon receptor (IFN-γR) impairs nerve injury-evoked activation of ipsilateral microglia and tactile allodynia. Reference Tsuda, Masuda, Kitano, Shimoyama, Tozaki-Saitoh and Inoue118 IFN-γ also increases dorsal horn excitability Reference Vikman, Duggan and Siddall119 and facilitates synaptic transmission between primary afferent C-fibres and Lamina 1 neurons via a microglial dependent mechanism. Reference Reischer, Heinke and Sandkuhler120

Failure to Resolve Chronic Neuroinflammation

All types of injury are capable of promoting inflammation and pain Reference Ji, Xu, Strichartz and Serhan121 and the interactions of inflammatory mediators with neurons, glia, immunocompetent leucocytes and lymphocytes, and macrophages Reference Scholz and Woolf14 promote neuroinflammation. Since identified “off signals” actively suppress the classical signs of inflammation, Reference Ji, Xu, Strichartz and Serhan121,Reference Ji122 pain is usually short lasting or acute. The signals that actively resolve inflammation and pain include anti-inflammatory cytokines such as IL-10 and lipid-derived specialized pro-resolving mediators (SPMs). Reference Buckley, Gilroy and Serhan123,Reference Buckley, Gilroy, Serhan, Stockinger and Tak124 Despite this, the neuroinflammation associated with neuropathic pain may not resolve, thereby promoting the transition from acute pain to chronic pain. Reference Ji, Xu, Strichartz and Serhan121 As already mentioned, spontaneous and ectopic activity in primary afferent fibers is crucial for the maintenance and persistence of signs of neuropathic pain. Reference Pitcher and Henry50Reference Yatziv and Devor53,Reference Haroutounian, Ford and Frey56 Excessive neuronal activity releases glutamate and neuropeptides which interact with glia and immune cells to provoke the generation of inflammatory mediators. Reference Xanthos and Sandkuhler73 It is possible that this incessant neurogenic neuroinflammation overcomes the resolution processes that normally terminate inflammation thereby contributing to the indefinite persistence of neuropathic pain.

In addition, the injury-induced structural changes in peripheral afferent Reference Gangadharan, Zheng and Tabener29 and sympathetic nerves Reference McLachlan, Janig and Michalis30,Reference Yen, Bennett and Ribeiro-da-Silva31 and in higher brain structures are almost certainly irreversible. Reference Price, Basbaum and Bresnahan12 These enduring changes also contribute to the chronic nature of neuropathic pain.

Changes in Central Sensory Pathways in Higher Brain Regions

Cytokine/chemokine/growth factor/glial cell interactions are also involved in modulation of sensory information in the mesolimbic system, Reference Taylor, Castonguay and Taylor125 thalamus, sensory cortex, nucleus accumbens, and amygdala. Reference Taylor, Castonguay and Taylor125Reference Wu and Zhu127 Peripheral nerve injury promotes microglial activation in the contralateral thalamus, sensory cortex, and amygdala as would be expected from the anatomical projections of ascending sensory fibers. Brain regions not directly involved in either sensory or affective aspects of pain, such as the motor cortex, do not display microglial activation. Reference Taylor, Mehrabani, Liu, Taylor and Cahill128 Hyperactivity in parts of the anterior cingulate cortex and other limbic structures drives the anxiety and depression that represent a co-morbidity of chronic and neuropathic pain. Reference Bannister, Sachau, Baron and Dickenson7,Reference Sellmeijer, Mathis and Hugel129

Blood-borne inflammatory mediators Reference Sandy-Hindmarch, Bennett, Wiberg, Furniss, Baskozos and Schmid130 from the site of peripheral injury increase the permeability of the blood–brain barrier. Reference Xanthos, Pungel, Wunderbaldinger and Sandkuhler33 This allows CNS neurons to access blood cells and the cytokines and chemokines they produce. Reference Greenhalgh, David and Bennett131 In addition, the selective activation of glia and immune cells in nociceptive pathways Reference Taylor, Castonguay and Taylor125 likely reflects localized neurogenic neuroinflammation in response to enduring intense activity. Reference Xanthos and Sandkuhler73

Alterations in Descending Control of Spinal Processing

Spinal nociceptive processing is subject to modulation by descending serotonergic and noradrenergic pathways. Reference Finnerup, Kuner and Jensen6,Reference Bannister and Dickenson132 Descending inhibition is mediated via α2-adrenoceptors and 5HT7 receptors whereas serotonergic activation of metabotropic 5HT2 receptors and ionotropic 5HT3 receptors facilitates transmission Reference Bannister, Sachau, Baron and Dickenson7 . Brainstem excitatory pathways are more important in the maintenance than in the induction of pain and under these conditions, α2-noradrenergic inhibition is attenuated whilst facilitation through 5HT2 and 5HT3 receptors is enhanced. Reference Bannister, Sachau, Baron and Dickenson7,Reference Bannister and Dickenson132Reference Bannister, Lockwood, Goncalves, Patel and Dickenson134 Actions on these descending controls are thus likely to underlie the efficacy of tricyclic antidepressants and serotonin-noradrenaline reuptake inhibitors in pain management. Reference Bannister, Sachau, Baron and Dickenson7,Reference Finnerup, Attal and Haroutounian10

Different Injuries and Different Etiologies

As already stated, different types of nerve injury provoke different types of behavioral or physiological response in both humans and animals. Reference Mogil1Reference Fitzgerald and McKelvey4 Thus while mechanical allodynia produced in animals by SNI Reference Decosterd and Woolf13 persists for many weeks, that produced by CCI is short-lived and recovery is seen in about 4 weeks. Reference Decosterd and Woolf13,Reference Noh, Mikler, Joy and Smith37 Similarly, changes in synaptic transmission in the superficial dorsal horn are more robust after sciatic CCI than after complete sciatic nerve section (axotomy). Reference Chen, Balasubramanyan, Lai, Todd and Smith92 These findings are consistent with the observation that CCI promotes stronger and more long-lasting upregulation of TNF-α, IL-1β, and CCL-2 than axotomy by nerve crush. Reference Kleinschnit, Brinkhoff, Zelenka, Sommer and Stoll135 It has also been shown that the neuronal subtypes in the dorsal horn that are involved in generation of mechanical allodynia is defined by the nature of peripheral nerve injury. Reference Peirs, Williams and Zhao136

More clinically relevant observations include reports that neuropathic pain associated with multiple sclerosis is characterized by loss of spinal neurons Reference Gushchina, Pryce and Yip137 but this effect is not seen with CCI. Reference Polgar, Hughes, Riddell, Maxwell, Puskár and Todd138,Reference Polgar, Gray, Riddell and Todd139 The above findings imply that different types of injury provoke the generation of different sets of mediators Reference Boakye, Tang and Smith18,Reference DeLeo, Colburn and Rickman140 and thus present different drug targets.

The Way Forward? Bridging the Gap between Basic Science and Clinical Practice

Given that patients with neuropathic pain are heterogeneous in pathophysiology, etiology, and clinical presentation Reference Mogil1,Reference Baron, Maier and Attal5 it is hardly surprising that injury-specific pathologies are found in animal models. As in the clinic, there is the added complication that signs of pain and response to medication of each experimental animal is determined by factors such as their sex, prior exposure to neonatal injury, age, intestinal microbiome, and environmental factors. Reference Mogil1Reference Fitzgerald and McKelvey4,Reference Brewer and Bacceic141,Reference Gaudet, Fonken, Ayala, Maier and Watkins142

Quantitative sensory testing (QST) may help to bridge the knowledge gap between clinical and laboratory findings. This involves formalization and quantification of a battery of neurological tests, such as response to von Frey filaments, vibration, heat, pressure, and cold as well as dynamic allodynia and wind-up ratio. Reference Baron, Maier and Attal5 Findings are compared with datasets that represent normal responses to sensory tests. Neuropathic pain patients can then be grouped into clusters based on their sensory profiles and this may have a role in determining treatment. Reference Vollert, Maier and Attal143 Technological improvements in microneurography have shown that the specific C-fibre subpopulation affected (mechanoinsensitive versus non-mechanoceptive) depends on the source of neuropathic pain and the type of neuropathy. Reference Serra, Bostock and Sola144,Reference Middleton, Barry and Comini145 Modern microneurography approaches will thus play a role in future refinement of QST. The validity of QST is supported by the observation that post hoc analysis of responders to treatments in clinical trials suggest that clinical effectiveness may cluster according to pain phenotype. Reference Vollert, Maier and Attal143 Beyond this, it may also be possible to subcategorize patients according to their cytokine profile. It then may be possible to correlate precisely quantified signs and symptoms in each individual patient to pathophysiology at the cellular and molecular level.

Recent improvements in basic science approaches also seek to bridge the gap between the “bench and bedside”. For example, improved methodologies are starting to differentiate probable pain in animal models from nociception or simple withdrawal reflexes. Reference Alles and Smith57,Reference Mogil146 Also more attention is now paid to the genetics, environment, and sex of experimental animals Reference Mogil1,Reference Mapplebeck, Beggs and Salter80 and improved methodologies are now available for bringing human tissue to the laboratory. These include the culture of human nociceptors either from surgical or post-mortem tissue or using human-induced pluripotent stem cell-derived nociceptors. Reference Middleton, Barry and Comini145,Reference Alsaloum and Waxman147

Taken together, these approaches will permit a rational and highly personalized medicine approach that will dictate the most appropriate therapeutic approach for each individual patient. Reference Bannister, Sachau, Baron and Dickenson7,Reference Renthal, Chamessian and Curatolo148,Reference Bouali-Benazzouz, Landry, Benazzouz and Fossat149

Funding

No financial support was provided for the writing of this review.

Disclosures

The author has no financial or other disclosures.

Statement of Authorship

PAS was responsible for conceiving, researching, and writing this article.

References

Mogil, JS. Sources of individual differences in pain. Ann Rev Neurosci. 2021;44:125. DOI 10.1523/JNEUROSCI.1786-18.CrossRefGoogle ScholarPubMed
Moriarty, O, Tu, Y, Sengar, AS, Salter, MW, Beggs, S, Walker, SM. Priming of adult incision response by early-life injury: neonatal microglial inhibition has persistent but sexually dimorphic effects in adult rats. J Neurosci. 2019;39:308193. DOI 10.1523/JNEUROSCI.1786-18.2019.CrossRefGoogle ScholarPubMed
Dworsky-Fried, Z, Kerr, BJ, Taylor, AMW. Microbes, microglia and pain. Neurobiol Pain. 2020;7:100045. DOI 10.1016/j.ynpai.2020.100045.CrossRefGoogle ScholarPubMed
Fitzgerald, M, McKelvey, R. Nerve injury and neuropathic pain - a question of age. Exp Neurol. 2016;275:296302. DOI 10.1016/j.expneurol.2015.07.013.CrossRefGoogle ScholarPubMed
Baron, R, Maier, C, Attal, N, et al. Peripheral neuropathic pain: a mechanism-related organizing principle based on sensory profiles. Pain. 2017;158:26172. DOI 10.1097/j.pain.0000000000000753.CrossRefGoogle ScholarPubMed
Finnerup, NB, Kuner, R, Jensen, JT. Neuropathic pain: from mechanisms to treatment. Physiol Rev. 2021;101:259301. DOI 10.1152/physrev.00045.2019.CrossRefGoogle ScholarPubMed
Bannister, K, Sachau, J, Baron, R, Dickenson, AH. Neuropathic pain: mechanism-based therapeutics. Ann Rev Pharmacol Toxicol. 2020;60:25774. DOI 10.1146/annurev-pharmtox-010818-021524.CrossRefGoogle ScholarPubMed
Wall, PD, Devor, M, Inbal, R, et al. Autotomy following peripheral nerve lesions: experimental anaesthesia dolorosa . Pain. 1979;7:10313. DOI 10.1016/0304-3959(79)90002-2.CrossRefGoogle ScholarPubMed
Alles, SRA, Smith, PA. The etiology and pharmacology of neuropathic pain. Pharmacol Rev. 2018;70:31547. DOI 10.1124/pr.117.014399.CrossRefGoogle ScholarPubMed
Finnerup, NB, Attal, N, Haroutounian, S, et al. Pharmacotherapy for neuropathic pain in adults: a systematic review and meta-analysis. Lancet Neurol. 2015;14:16273. DOI 10.1016/S1474-4422(14)70251-0.CrossRefGoogle ScholarPubMed
Gormsen, L, Rosenberg, R, Bach, FW, Jensen, TS. Depression anxiety health-related quality of life and pain in patients with chronic fibromyalgia and neuropathic pain. Eur J Pain. 2010;14:1278. DOI 10.1016/j.ejpain.2009.03.010.CrossRefGoogle ScholarPubMed
Price, TJ, Basbaum, AI, Bresnahan, J, et al. Transition to chronic pain: opportunities for novel therapeutics. Nat Rev Neurosci. 2018;19:3834. DOI 10.1038/s41583-018-0012-5.CrossRefGoogle ScholarPubMed
Decosterd, I, Woolf, CJ. Spared nerve injury: an animal model of persistent peripheral neuropathic pain. Pain. 2000;87:14958. DOI 10.1016/S0304-3959(00)00276-1.CrossRefGoogle ScholarPubMed
Scholz, J, Woolf, CJ. The neuropathic pain triad: neurons immune cells and glia. Nat Neurosci. 2007;10:13618. DOI 10.1038/nn1992.CrossRefGoogle ScholarPubMed
Moalem, G, Tracey, DJ. Immune and inflammatory mechanisms in neuropathic pain. Brain Res Rev. 2006;51:24064. DOI 10.1016/j.brainresrev.2005.11.004.CrossRefGoogle ScholarPubMed
Marais, R, Light, Y, Mason, C, Paterson, H, Olson, MF, Marshall, CJ. Requirement of Ras-GTP-Raf complexes for activation of Raf-1 by protein kinase C. Science. 1998;280:10912. DOI 10.1126/science.280.5360.109.CrossRefGoogle ScholarPubMed
Leung, L, Cahill, CM. TNF-alpha and neuropathic pain–a review. J Neuroinflammation. 2010;7:27. DOI 10.1186/1742-2094-7-27.CrossRefGoogle ScholarPubMed
Boakye, PA, Tang, SJ, Smith, PA. Mediators of neuropathic pain; focus on spinal microglia, CSF-1, BDNF, CCL21, TNF-alpha, Wnt ligands and interleukin 1-beta. Front Pain Res. 2021;2:41. DOI 10.3389/fpain.2021.698157.CrossRefGoogle Scholar
Vasudeva, K, Vodovotz, Y, Azhar, N, Barclay, D, Janjic, JM, Pollock, JA. In vivo and systems biology studies implicate IL-18 as a central mediator in chronic pain. J Neuroimmunol. 2015;283:439. DOI 10.1016/j.jneuroim.2015.04.012.CrossRefGoogle ScholarPubMed
Gomez-Nicola, D, Valle-Argos, B, Suardiaz, M, Taylor, JS, Nieto-Sampedro, M. Role of IL-15 in spinal cord and sciatic nerve after chronic constriction injury: regulation of macrophage and T-cell infiltration. J Neurochem. 2008;107:174152. DOI 10.1111/j.1471-4159.2008.05746.x.CrossRefGoogle ScholarPubMed
Kleinschnitz, C, Hofstetter, HH, Meuth, SG, Braeuninger, S, Sommer, C, Stoll, G. T cell infiltration after chronic constriction injury of mouse sciatic nerve is associated with interleukin-17 expression. Exp Neurol. 2006;200:4805. DOI 10.1016/j.expneurol.2006.03.014.CrossRefGoogle ScholarPubMed
Pezet, S, McMahon, SB. Neurotrophins: mediators and modulators of pain. Ann Rev Neurosci. 2006;29:50738. DOI 10.1146/annurev.neuro.29.051605.112929.CrossRefGoogle ScholarPubMed
White, FA, Wilson, NM. Chemokines as pain mediators and modulators. Curr Opin Anaesthesiol. 2008;21:5805. DOI 10.1097/ACO.0b013e32830eb69d.CrossRefGoogle ScholarPubMed
Silva, RL, Lopes, AH, Guimaraes, RM, Cunha, TM. CXCL1/CXCR2 signaling in pathological pain: role in peripheral and central sensitization. Neurobiol Dis. 2017;105:10916. DOI 10.1016/j.nbd.2017.06.001.CrossRefGoogle ScholarPubMed
Yu, Y, Huang, X, Di, Y, Qu, L, Fan, N. Effect of CXCL12/CXCR4 signaling on neuropathic pain after chronic compression of dorsal root ganglion. Sci Rep. 2017;7:5707. DOI 10.1038/s41598-017-05954-1.CrossRefGoogle ScholarPubMed
Kaur, G, Singh, N, Jaggi, AS. Mast cells in neuropathic pain: an increasing spectrum of their involvement in pathophysiology. Rev Neurosci. 2017;28:75966. DOI 10.1515/revneuro-2017-0007.CrossRefGoogle ScholarPubMed
ObaraI Telezhkin, V, Alrashdi, I, Chazot, PL. Histamine histamine receptors and neuropathic pain relief. Br J Pharmacol. 2020;177:58099. DOI 10.1111/bph.14696.Google Scholar
van Vliet, AC, Lee, J, van der Poel, M, et al. Coordinated changes in the expression of Wnt pathway genes following human and rat peripheral nerve injury. PLoS One. 2021;16:e0249748. DOI 10.1371/journal.pone.0249748.CrossRefGoogle ScholarPubMed
Gangadharan, V, Zheng, H, Tabener, FJ, et al. Neuropathic pain caused by miswiring and abnormal end organ targeting. Nature. 2022;606:13745. DOI 10.1038/s41586-022-04777-z.CrossRefGoogle ScholarPubMed
McLachlan, EM, Janig, W, Michalis, M. Peripheral nerve injury triggers noradrenergic sprouting within dorsal root ganglia. Nature. 1993;363:5436. DOI 10.1038/363543a0.CrossRefGoogle ScholarPubMed
Yen, LD, Bennett, GJ, Ribeiro-da-Silva, A. Sympathetic sprouting and changes in nociceptive sensory innervation in the glabrous skin of the rat hind paw following partial peripheral nerve injury. J Comp Neurol. 2006;495:67990. DOI 10.1002/cne.20899.CrossRefGoogle ScholarPubMed
Drummond, PD, Drummond, SD, Dawson, LF, et al. Upregulation of alpha-1 adrenoceptors on cutaneous nerve fibres after partial sciatic nerve ligation and in complex regional pain syndrome type II. Pain. 2014;155:60616. DOI 10.1016/j.pain.2013.12.021.CrossRefGoogle ScholarPubMed
Xanthos, DN, Pungel, I, Wunderbaldinger, G, Sandkuhler, J. Effects of peripheral inflammation on the blood-spinal cord barrier. Mol Pain. 2012;8:44. DOI 10.1186/1744-8069-8-44.CrossRefGoogle ScholarPubMed
Lim, TKY, Shi, XQ, Martin, HC, et al. Blood-nerve barrier dysfunction contributes to the generation of neuropathic pain and allows targeting of injured nerves for pain relief. Pain. 2014;155:95467. DOI 10.1016/j.pain.2014.01.026.CrossRefGoogle Scholar
DiSabato, DJ, Quan, N, Godbout, JP. Neuroinflammation: the devil is in the details. J Neurochem. 2016;139:13653. DOI 10.1111/jnc.13607.CrossRefGoogle ScholarPubMed
Milatovic, D, Zaja-Milatovic, S, Breyer, RM, Aschner, M, Montine, TJ. Neuroinflammation and oxidative injury. In: Gupta, RC, editor. Reproductive and developmental toxicology. Cambridge: Academic Press; 2017, pp. 10511061.Google Scholar
Noh, MC, Mikler, B, Joy, T, Smith, PA. Time course of inflammation in dorsal root ganglia correlates with differential reversibility of mechanical allodynia. Neuroscience. 2020;428:199216. DOI 10.1016/j.neuroscience.2019.12.040.CrossRefGoogle Scholar
Yuan, Q, Liu, X, Xian, Y-F, et al. Satellite glia activation in dorsal root ganglion contributes to mechanical allodynia after selective motor fiber injury in adult rats. Biomed Pharmacother. 2020;127:110187. DOI 10.1016/j.biopha.2020.110187.CrossRefGoogle ScholarPubMed
Xie, W, Strong, JA, Zhang, JM. Early blockade of injured primary sensory afferents reduces glial cell activation in two rat neuropathic pain models. Neuroscience. 2009;160:84757. DOI 10.1016/j.neuroscience.2009.03.016.CrossRefGoogle ScholarPubMed
Zhang, X, Xiao, HS. Gene array analysis to determine the components of neuropathic pain signaling. Curr Opin Mol Ther. 2005;7:5327.Google ScholarPubMed
Biber, K, Boddeke, E. Neuronal CC chemokines: the distinct roles of CCL21 and CCL2 in neuropathic pain. Front Cell Neurosci. 2014;8:210. DOI 10.3389/fncel.2014.00210.CrossRefGoogle ScholarPubMed
Baskozos, GJ. Comprehensive analysis of long noncoding RNA expression in dorsal root ganglion reveals cell-type specificity and dysregulation after nerve injury. Pain. 2019;160:46385. DOI 10.1097/j.pain.0000000000001416.CrossRefGoogle ScholarPubMed
Liu, L, Xu, D, Wang, T, et al. Epigenetic reduction of miR-214-3p upregulates astrocytic colony-stimulating factor-1 and contributes to neuropathic pain induced by nerve injury. Pain. 2020;161:96108. DOI 10.1097/j.pain.0000000000001681.CrossRefGoogle ScholarPubMed
Zhang, YU, Ye, G, Zhao, J, et al. Exosomes carried miR-181c-5p alleviates neuropathic pain in CCI rat models. An Acad Bras Cienc. 2022;94:e20210564. DOI 10.1590/0001-3765202220210564.CrossRefGoogle ScholarPubMed
Noh, MC, Stemkowski, PL, Smith, PA. Long-term actions of interleukin-1beta on K(+) Na(+) and Ca(2+) channel currents in small IB4-positive dorsal root ganglion neurons; possible relevance to the etiology of neuropathic pain. J Neuroimmunol. 2019;332:198211. DOI 10.1016/j.jneuroim.2019.05.002.CrossRefGoogle Scholar
Stemkowski, PL, Smith, PA. Long-term IL-1beta exposure causes subpopulation-dependent alterations in rat dorsal root ganglion neuron excitability. J Neurophysiol. 2012;107:158697. DOI 10.1152/jn.00587.2011.CrossRefGoogle ScholarPubMed
Stemkowski, PL, Noh, MC, Chen, Y, Smith, PA. Increased excitability of medium-sized dorsal root ganglion neurons by prolonged interleukin-1beta exposure is K+ channel dependent and reversible. J Physiol. 2015;593:373955. DOI 10.1113/JP270905.CrossRefGoogle Scholar
Binshtok, AM, Wang, H, Zimmermann, K, et al. Nociceptors are interleukin-1{beta} sensors. J Neurosci. 2008;28:1406273. DOI 10.1523/JNEUROSCI.3795-08.2008.CrossRefGoogle ScholarPubMed
Chen, X, Pang, RP, Shen, KF, et al. TNF-alpha enhances the currents of voltage gated sodium channels in uninjured dorsal root ganglion neurons following motor nerve injury. Exp Neurol. 2011;227:27986. DOI 10.1016/j.expneurol.2010.11.017.CrossRefGoogle ScholarPubMed
Pitcher, GM, Henry, JL. Governing role of primary afferent drive in increased excitation of spinal nociceptive neurons in a model of sciatic neuropathy. Exp Neurol. 2008;214:21928. DOI 10.1016/j.expneurol.2008.08.003.CrossRefGoogle Scholar
Haroutounian, S, Nikolajsen, L, Bendtsen, TF, et al. Primary afferent input critical for maintaining spontaneous pain in peripheral neuropathy. Pain. 2014;155:12729. DOI 10.1016/j.pain.2014.03.022.CrossRefGoogle ScholarPubMed
Vaso A.Adahan, HM, Gjika, A, et al. Peripheral nervous system origin of phantom limb pain. Pain. 2014;155:138491. DOI 10.1016/j.pain.2014.04.018.Google Scholar
Yatziv, SL, Devor, M. Suppression of neuropathic pain by selective silencing of dorsal root ganglion ectopia using nonblocking concentrations of lidocaine. Pain. 2019;160:210514. DOI 10.1097/j.pain.0000000000001602.CrossRefGoogle ScholarPubMed
Dworkin, RH, O’Connor, AB, Backonja, M, et al. Pharmacologic management of neuropathic pain: evidence-based recommendations. Pain. 2007;132:23751. DOI 10.1016/j.pain.2007.08.033.CrossRefGoogle ScholarPubMed
Ritter, DM, Zemel, BM, Hala, TJ, O’Leary, ME, Lepore, AC, Covarrubais, M. Dysregulation of Kv3.4 channels in dorsal root ganglia following spinal cord injury. J Neurosci. 2015;35:126073. DOI 10.1523/JNEUROSCI.1594-14.2015.CrossRefGoogle ScholarPubMed
Haroutounian, S, Ford, AL, Frey, K, et al. How central is central poststroke pain? The role of afferent input in poststroke neuropathic pain: a prospective open-label pilot study. Pain. 2018;159:131724. DOI 10.1097/j.pain.0000000000001213.CrossRefGoogle ScholarPubMed
Alles, SRA, Smith, PA. Peripheral voltage-gated cation channels in neuropathic pain and their potential as therapeutic targets. Front Pain Res. 2021;2:750583. DOI 10.3389/fpain.2021.750583.10.3389/fpain.2021.750583CrossRefGoogle ScholarPubMed
Eschenfelder, S, Habler, HJ, Jannig, W. Dorsal root section elicits signs of neuropathic pain rather than reversing them in rats with L5 spinal nerve injury. Pain. 2000;87:2139. DOI 10.1016/S0304-3959(00)00285-2.CrossRefGoogle ScholarPubMed
Stemkowski, PL, Smith, PA. Sensory neurons, ion channels inflammation and the onset of neuropathic pain. Can J Neurol Sci. 2012;39:41635. DOI 10.1017/s0317167100013937.CrossRefGoogle ScholarPubMed
Wolf, G, Gabay, E, Tal, M, Yirmiya, R, Shavit, Y. Genetic impairment of interleukin-1 signaling attenuates neuropathic pain autotomy and spontaneous ectopic neuronal activity following nerve injury in mice. Pain. 2006;120:31524. DOI 10.1016/j.pain.2005.11.011.CrossRefGoogle ScholarPubMed
Andrade P.Hoogland, G, Del Rosario, J, et al. Tumor necrosis factor-alpha inhibitors alleviation of experimentally induced neuropathic pain is associated with modulation of TNF receptor expression. J Neurosci Res. 2014;92:14908. DOI 10.1002/jnr.23432.Google Scholar
Al-Mazidi, S, Alotaibi, M, Nedjadi, T, Chaudhary, A, Alzoghaibi, M, Djouhri, L. Blocking of cytokines signalling attenuates evoked and spontaneous neuropathic pain behaviours in the paclitaxel rat model of chemotherapy-induced neuropathy. Eur J Pain. 2018;22:81021. DOI 10.1002/ejp.1169.CrossRefGoogle ScholarPubMed
Grace, PM, Hutchinson, MR, Maier, SF, Watkins, LR. Pathological pain and the neuroimmune interface. Nat Rev Immunol. 2014;14:21731. DOI 10.1038/nri3621.CrossRefGoogle ScholarPubMed
Yekkirala, AS, Roberson, DP, Bean, BP, Woolf, CJ. Breaking barriers to novel analgesic drug development. Nat Rev Drug Discov. 2017;16:54564. DOI 10.1038/nrd.2017.87.Google ScholarPubMed
Gustafson-Vickers, SL, Lu, VB, Lai, AY, Todd, KG, Ballanyi, K, Smith, PA. Long-term actions of interleukin-1beta on delay and tonic firing neurons in rat superficial dorsal horn and their relevance to central sensitization. Mol Pain. 2008;4:63. DOI 10.1186/1744-8069-4-63.CrossRefGoogle ScholarPubMed
Kawasaki, Y, Zhang, L, Cheng, JK, Ji, RR. Cytokine mechanisms of central sensitization: distinct and overlapping role of interleukin-1beta interleukin-6 and tumor necrosis factor-alpha in regulating synaptic and neuronal activity in the superficial spinal cord. J Neurosci. 2008;28:518994. DOI 10.1523/JNEUROSCI.3338-07.2008.CrossRefGoogle ScholarPubMed
Vikman, KS, Siddall, PJ, Duggan, AW. Increased responsiveness of rat dorsal horn neurons in vivo following prolonged intrathecal exposure to interferon-[gamma]. Neuroscience. 2005;135:96977. DOI 10.1016/j.pain.2007.02.010.CrossRefGoogle ScholarPubMed
Talbot, S, Foster, SL, Woolf, CJ. Neuroimmunity: physiology and pathology. Annu Rev Immunol. 2016;34:42147. DOI 10.1146/annurev-immunol-041015-055340.CrossRefGoogle ScholarPubMed
Pinho-Ribeiro, FA, Verri, WA Jr, Chiu, IM. Nociceptor sensory neuron-immune interactions in pain and inflammation. Trends Immunol. 2017;38:519. DOI 10.1016/j.it.2016.10.001.CrossRefGoogle ScholarPubMed
Chiu, IM, von Hehn, CA, Woolf, CJ. Neurogenic inflammation and the peripheral nervous system in host defense and immunopathology. Nat Neurosci. 2012;15:10637. DOI 10.1038/nn.3144.CrossRefGoogle ScholarPubMed
Schaible, HG, Del, RA, Matucci-Cerinic, M. Neurogenic aspects of inflammation. Rheum Dis Clin North Am. 2005;31:77101. DOI 10.1016/j.rdc.2004.09.004.CrossRefGoogle ScholarPubMed
McMahon, SB, La Russa, F, Bennett, DL. Crosstalk between the nociceptive and immune systems in host defence and disease. Nat Rev Neurosci. 2015;16:389402. DOI 10.1038/nrn3946.CrossRefGoogle ScholarPubMed
Xanthos, DN, Sandkuhler, J. Neurogenic neuroinflammation: inflammatory CNS reactions in response to neuronal activity. Nat Rev Neurosci. 2014;15:4353. DOI 10.1038/nrn3617.CrossRefGoogle ScholarPubMed
Shi, X, Wang, L, Li, X, Peymen, S, Kingerly, WS, Clark, JD. Neuropeptides contribute to peripheral nociceptive sensitization by regulating interleukin-1beta production in keratinocytes. Anesth Analg. 2011;113:17583. DOI 10.1213/ANE.0b013e31821a0258.CrossRefGoogle ScholarPubMed
Okubo, M, Yamanaka, H, Kobayashi, K, et al. Macrophage-colony stimulating factor derived from injured primary afferent induces proliferation of spinal microglia and neuropathic pain in rats. PLoS One. 2016;11:e0153375. DOI 10.1371/journal.pone.0153375.CrossRefGoogle ScholarPubMed
Guan, Z, Kuhn, JA, Wang, X, et al. Injured sensory neuron-derived CSF1 induces microglial proliferation and DAP12-dependent pain. Nat Neurosci. 2016;19:94101. DOI 10.1038/nn.4189.CrossRefGoogle ScholarPubMed
Malcangio, M. Role of the immune system in neuropathic pain. Scand J Pain. 2020;20:3337. DOI 10.1515/sjpain-2019-0138.CrossRefGoogle Scholar
Dong, J, Xu, C, Xia, R, Zhang, Z. Upregulating miR-130a-5p relieves astrocyte over activation-induced neuropathic pain through targeting C-X-C motif chemokine receptor 12/C-X-C motif chemokine receptor 4 axis. NeuroReport. 2021;32:13543. DOI 10.1097/WNR.0000000000001573.CrossRefGoogle ScholarPubMed
Halievski, K, Ghazisaeidi, S, Salter, MW. Sex-dependent mechanisms of chronic pain: a focus on microglia and P2X4R. J Pharmacol Exp Ther. 2020;375:2029. DOI 10.1124/jpet.120.265017.CrossRefGoogle ScholarPubMed
Mapplebeck, JC, Beggs, S, Salter, MW. Molecules in pain and sex: a developing story. Mol Brain. 2017;10:9. DOI 10.1186/s13041-017-0289-8.CrossRefGoogle ScholarPubMed
Sorge, RE, Mapplebeck, JC. Different immune cells mediate mechanical pain hypersensitivity in male and female mice. Nat Neurosci. 2015;18:10813. DOI 10.1038/nn.4053.CrossRefGoogle ScholarPubMed
van Weering, HR, de Jong, AP, de Haas, AH, Biber, KP, Boddeke, HW. CCL21-induced calcium transients and proliferation in primary mouse astrocytes: CXCR3-dependent and independent responses. Brain Behav Immun. 2010;24:76875. DOI 10.1016/j.bbi.2009.04.007.Google ScholarPubMed
Costigan, M, Moss, A, Latremoliere, A, et al. T-cell infiltration and signaling in the adult dorsal spinal cord is a major contributor to neuropathic pain-like hypersensitivity. J Neurosci. 2009;29:1441522. DOI 10.1523/JNEUROSCI.4569-09.2009.CrossRefGoogle Scholar
Biggs, JE, Lu, VB, Stebbing, MJ, Balasubramanyan, S, Smith, PA. Is BDNF sufficient for information transfer between microglia and dorsal horn neurons during the onset of central sensitization? Mol Pain. 2010;6:44. DOI 10.1186/1744-8069-6-44.CrossRefGoogle ScholarPubMed
Smith, PA. BDNF: no gain without pain? Neuroscience. 2014;283:10723. DOI 10.1016/j.neuroscience.2014.05.044.CrossRefGoogle ScholarPubMed
Yu, X, Basbaum, A, Guan, Z. Contribution of colony-stimulating factor 1 to neuropathic pain. PAIN Rep. 2021;6:e883. DOI 10.1097/PR9.0000000000000883.CrossRefGoogle ScholarPubMed
Boakye, PA, Rancic, V, Whitlock, KH, et al. Receptor dependence of BDNF actions in superficial dorsal horn: relation to central sensitization and actions of macrophage colony stimulating factor 1. J Neurophysiol. 2019;121:230822. DOI 10.1152/jn.00839.2018.CrossRefGoogle ScholarPubMed
Zhang W.Shi, Y, Peng, Y, et al. Neuron activity-induced Wnt signaling up-regulates expression of brain-derived neurotrophic factor in the pain neural circuit. J Biol Chem. 2018;293:1564151. DOI 10.1074/jbc.RA118.002840.Google Scholar
Trang, T, Beggs, S, Wan, X, Salter, MW. P2X4-receptor-mediated synthesis and release of brain-derived neurotrophic factor in microglia is dependent on calcium and p38-mitogen-activated protein kinase activation. J Neurosci. 2009;29:351828. DOI 10.1523/JNEUROSCI.5714-08.2009.CrossRefGoogle ScholarPubMed
Coull, JA, Boudreau, D, Bachand, K, et al. Trans-synaptic shift in anion gradient in spinal lamina I neurons as a mechanism of neuropathic pain. Nature. 2003;424:93842. DOI 10.1038/nature01868.CrossRefGoogle ScholarPubMed
Balasubramanyan, S, Stemkowski, PL, Stebbing, MJ, Smith, PA. Sciatic chronic constriction injury produces cell-type specific changes in the electrophysiological properties of rat substantia gelatinosa neurons. J Neurophysiol. 2006;96:57990. DOI 10.1152/jn.00087.2006.CrossRefGoogle ScholarPubMed
Chen, Y, Balasubramanyan, S, Lai, AY, Todd, KG, Smith, PA. Effects of sciatic nerve axotomy on excitatory synaptic transmission in rat substantia gelatinosa. J Neurophysiol. 2009;102:320315. DOI 10.1152/jn.00087.2006.CrossRefGoogle ScholarPubMed
Lu, VB, Ballanyi K.Colmers, WF, Smith, PA. Neuron type-specific effects of brain-derived neurotrophic factor in rat superficial dorsal horn and their relevance to ‘central sensitization’. J Physiol. 2007;584:54363. DOI 10.1113/jphysiol.2007.141267.CrossRefGoogle ScholarPubMed
Lu, VB, Biggs, JE, Stebbing, MJ, et al. BDNF drives the changes in excitatory synaptic transmission in the rat superficial dorsal horn that follow sciatic nerve injury. J Physiol. 2009;587:101332. DOI 10.1113/jphysiol.2008.166306.CrossRefGoogle ScholarPubMed
Hildebrand, ME, Jian, X, Dedek, A, et al. Potentiation of synaptic GluN2B NMDAR currents by fyn kinase is gated through BDNF-mediated disinhibition in spinal pain processing. Cell Rep. 2016;17:275365. DOI 10.1016/j.celrep.2016.11.024.CrossRefGoogle ScholarPubMed
Coull, JA, Beggs, S, Boudreau, D, et al. BDNF from microglia causes the shift in neuronal anion gradient underlying neuropathic pain. Nature. 2005;438:101721. DOI 10.1038/nature04223.CrossRefGoogle ScholarPubMed
Peirs, C, Seal, RP. Neural circuits for pain: recent advances and current views. Science. 2016;354:57884. DOI 10.1126/science.aaf8933.CrossRefGoogle ScholarPubMed
Peirs, C, Williams, SP, Zhao, X, et al. Dorsal horn circuits for persistent mechanical pain. Neuron. 2015;87:797812. DOI 10.1016/j.neuron.2015.07.029.CrossRefGoogle ScholarPubMed
Baba, H, Ji, R-R, Kohno, T, et al. Removal of GABAergic inhibition facilitates polysynaptic A fiber-mediated excitatory transmission to the superficial spinal dorsal horn. Mol Cell Neurosci. 2003;24:81830. DOI 10.1016/s1044-7431(03)00236-7.CrossRefGoogle Scholar
Prescott, SA, Ma, Q, De Koninck, Y. Normal and abnormal coding of somatosensory stimuli causing pain. Nat Neurosci. 2014;17:18391. DOI 10.1038/nn.3629.CrossRefGoogle ScholarPubMed
Kerr, BJ, Bradbury, EJ, Bennett, DL, et al. Brain-derived neurotrophic factor modulates nociceptive sensory inputs and NMDA-evoked responses in the rat spinal cord. J Neurosci. 1999;19:513848. DOI 10.1523/JNEUROSCI.19-12-05138.1999.CrossRefGoogle ScholarPubMed
Chen, W, Walwyn, W, Ennes, HS, et al. BDNF released during neuropathic pain potentiates NMDA receptors in primary afferent terminals. Eur J Neurosci. 2014;39:143954. DOI 10.1111/ejn.12516.CrossRefGoogle ScholarPubMed
Ferrini, F, Perez-Sanchez, J, Ferland, S, et al. Differential chloride homeostasis in the spinal dorsal horn locally shapes synaptic metaplasticity and modality-specific sensitization. Nat Commun. 2020;11:3935. DOI 10.1038/s41467-020-17824-y.CrossRefGoogle ScholarPubMed
Ferrini, F, De Koninck, Y. Microglia control neuronal network excitability via BDNF signalling. Neural Plast. 2013;2013:429815. DOI 10.1155/2013/429815.CrossRefGoogle ScholarPubMed
Sandkuhler, J, Benrath, J, Brechtel, C, Ruscheweyh, R, Heinke, B. Synaptic mechanisms of hyperalgesia. Prog Brain Res. 2000;129:81100. DOI 10.1016/S0079-6123(00)29007-9.CrossRefGoogle ScholarPubMed
Sandkuhler, J. Understanding LTP in pain pathways. Mol Pain. 2007;3:9. DOI 10.1016/S0079-6123(00)29007-9.CrossRefGoogle ScholarPubMed
Li, S, Cai, J. BDNF contributes to spinal long-term potentiation and mechanical hypersensitivity via fyn-mediated phosphorylation of NMDA receptor GluN2B subunit at Tyrosine 1472 in rats following spinal nerve ligation. Neurochem Res. 2017;42:271229. DOI 10.1007/s11064-017-2274-0.CrossRefGoogle ScholarPubMed
Ding, X, Cia, J, li, S, et al. BDNF contributes to the development of neuropathic pain by induction of spinal long-term potentiation via SHP2 associated GluN2B-containing NMDA receptors activation in rats with spinal nerve ligation. Neurobiol Dis. 2015;73:42851. DOI 10.1016/j.nbd.2014.10.025.CrossRefGoogle Scholar
Zhou, LJ, Peng, J, Xu, YN, et al. Microglia are indispensable for synaptic plasticity in the spinal dorsal horn and chronic pain. Cell Rep. 2019;27:384459. DOI 10.1016/j.celrep.2019.05.087.CrossRefGoogle ScholarPubMed
Mapplebeck, JCS, Lorenzo, LE, Lee, KY, et al. Chloride dysregulation through downregulation of KCC2 mediates neuropathic pain in both sexes. Cell Rep. 2019;28:5906. DOI 10.1016/j.celrep.2019.06.059.CrossRefGoogle ScholarPubMed
Paige, C, Plasencia-Fernandez, I, Kume, M, et al. A female-specific role for calcitonin gene-related peptide (CGRP) in rodent pain models. J Neurosci. 2022;42:193044. DOI 10.1523/JNEUROSCI.1137-21.2022.CrossRefGoogle ScholarPubMed
Gardell, LR, Vanderah, TW, Gardell, SE, et al. Enhanced evoked excitatory transmitter release in experimental neuropathy requires descending facilitation. J Neurosci. 2002;23:83709. DOI 10.1523/JNEUROSCI.23-23-08370.2003.Google Scholar
Gajtko, A, Bakk, E, Hegedus, K, Ducza, L, Hollo, K. IL-1beta induced cytokine expression by spinal astrocytes can play a role in the maintenance of chronic inflammatory pain. Front Physiol. 2020;11:543331. DOI 10.3389/fphys.2020.543331.CrossRefGoogle ScholarPubMed
Yan, X, Maixner, Li F, etal, DW. Interleukin-1beta released by microglia initiates the enhanced glutamatergic activity in the spinal dorsal horn during paclitaxel-associated acute pain syndrome. Glia. 2019;67:48297. DOI 10.1002/glia.23557.CrossRefGoogle ScholarPubMed
Gruber-Schoffnegger, D, Drdla-Schutting, R, Honigsperger, C, Wunderbaldinger, G, Gassner, M, Sandkuhler, J. Induction of thermal hyperalgesia and synaptic long-term potentiation in the spinal cord lamina I by TNF-alpha and IL-1beta is mediated by glial cells. J Neurosci. 2013;33:654051. DOI 10.1523/JNEUROSCI.5087-12.2013.CrossRefGoogle ScholarPubMed
del Rivero, T, Fischer, R, Yang, F, Swanson, KA, Bethea, JR. Tumor necrosis factor receptor 1 inhibition is therapeutic for neuropathic pain in males but not in females. Pain. 2019;160:92231. DOI 10.1097/j.pain.0000000000001470.CrossRefGoogle Scholar
Goncalves Dos, SG, Delay, L, Yaksh, TL, Corr, M. Neuraxial cytokines in pain states. Front Immunol. 2020;10:3061. DOI 10.3389/fimmu.2019.03061.Google Scholar
Tsuda, M, Masuda, T, Kitano, J, Shimoyama, H, Tozaki-Saitoh, H, Inoue, K. IFN-gamma receptor signaling mediates spinal microglia activation driving neuropathic pain. Proc Natl Acad Sci U S A. 2009;106:80327. DOI 10.1073/pnas.0810420106.CrossRefGoogle ScholarPubMed
Vikman, KS, Duggan, AW, Siddall, PJ. Interferon-gamma induced disruption of GABAergic inhibition in the spinal dorsal horn in vivo. Pain. 2007;133:1828. DOI 10.1016/j.pain.2007.02.010.CrossRefGoogle ScholarPubMed
Reischer, G, Heinke, B, Sandkuhler, J. Interferon gamma facilitates the synaptic transmission between primary afferent C-fibres and lamina I neurons in the rat spinal dorsal horn via microglia activation. Mol Pain. 2020;16:1744806920917249. DOI 10.1177/1744806920917249.CrossRefGoogle ScholarPubMed
Ji, RR, Xu, ZZ, Strichartz, G, Serhan, CN. Emerging roles of resolvins in the resolution of inflammation and pain. Trends Neurosci. 2011;34:599609. DOI 10.1016/j.tins.2011.08.005.CrossRefGoogle ScholarPubMed
Ji, RR. Specialized pro-resolving mediators as resolution pharmacology for the control of pain and itch. Annu Rev Pharmacol Toxicol. 2022;63:27393. DOI 10.1146/annurev-pharmtox-051921-084047.Google ScholarPubMed
Buckley, CD, Gilroy, DW, Serhan, CN. Proresolving lipid mediators and mechanisms in the resolution of acute inflammation. Immunity. 2014;40:31527. DOI 10.1016/j.immuni.2014.02.009.CrossRefGoogle ScholarPubMed
Buckley, CD, Gilroy, DW, Serhan, CN, Stockinger, B, Tak, PP. The resolution of inflammation. Nat Rev Immunol. 2013;13:5966. DOI 10.1038/nri3362.CrossRefGoogle ScholarPubMed
Taylor, AMW, Castonguay, A, Taylor, AJ, et al. Microglia disrupt mesolimbic reward circuitry in chronic pain. J Neurosci. 2015;35:844250. DOI 10.1523/JNEUROSCI.4036-14.2015.CrossRefGoogle ScholarPubMed
Fiore, NT, Austin, PJ. Peripheral nerve injury triggers neuroinflammation in the medial prefrontal cortex and ventral hippocampus in a subgroup of rats with coincident affective behavioural changes. Neuroscience. 2019;416:14767. DOI 10.1016/j.neuroscience.2019.08.005.CrossRefGoogle Scholar
Wu, XB, Zhu, Q. GaoYJ CCL2/CCR2 contributes to the altered excitatory-inhibitory synaptic balance in the nucleus accumbens shell following peripheral nerve injury-induced neuropathic pain. Neurosci Bull. 2021;37:92133. DOI 10.1007/s12264-021-00697-6.CrossRefGoogle Scholar
Taylor, AMW, Mehrabani, S, Liu, S, Taylor, AJ, Cahill, CM. Topography of microglial activation in sensory- and affect-related brain regions in chronic pain. J Neurosci Res. 2017;95:13305. DOI 10.1002/jnr.23883.CrossRefGoogle ScholarPubMed
Sellmeijer, J, Mathis, V, Hugel, S, et al. Hyperactivity of anterior cingulate cortex areas 24a/24b drives chronic pain-induced anxiodepressive-like consequences. J Neurosci. 2018;38:310215. DOI 10.1523/JNEUROSCI.3195-17.2018.CrossRefGoogle ScholarPubMed
Sandy-Hindmarch, O, Bennett, DL, Wiberg, A, Furniss, D, Baskozos, G, Schmid, A. Systemic inflammatory markers in neuropathic pain nerve injury and recovery. Pain. 2022;163:52637.CrossRefGoogle ScholarPubMed
Greenhalgh, AD, David, S, Bennett, FC. Immune cell regulation of glia during CNS injury and disease. Nat Rev Neurosci. 2020;21:13952. DOI 10.1038/s41583-020-0263-9.CrossRefGoogle ScholarPubMed
Bannister, K, Dickenson, AH. The plasticity of descending controls in pain: translational probing. J Physiol. 2017;595:415966. DOI 10.1113/JP274165.CrossRefGoogle ScholarPubMed
Bannister, K, Patel, R, Goncalves, L, Townson, L, Dickenson, AH. Diffuse noxious inhibitory controls and nerve injury: restoring an imbalance between descending monoamine inhibitions and facilitations. Pain. 2015;156:180311. DOI 10.1097/j.pain.0000000000000240.CrossRefGoogle ScholarPubMed
Bannister, K, Lockwood, S, Goncalves, L, Patel, R, Dickenson, AH. An investigation into the inhibitory function of serotonin in diffuse noxious inhibitory controls in the neuropathic rat. Eur J Pain. 2017;21:75060. DOI 10.1002/ejp.979.CrossRefGoogle ScholarPubMed
Kleinschnit, C, Brinkhoff, J, Zelenka, M, Sommer, C, Stoll, G. The extent of cytokine induction in peripheral nerve lesions depends on the mode of injury and NMDA receptor signaling. J Neuroimmunol. 2004;149:7783. DOI 10.1016/j.jneuroim.2003.12.013.Google Scholar
Peirs, C, Williams, S-PG, Zhao, X, et al. Mechanical allodynia circuitry in the dorsal horn is defined by the nature of the injury. Neuron. 2021;109:7390. DOI 10.1016/j.neuron.2020.10.027.CrossRefGoogle ScholarPubMed
Gushchina, S, Pryce, G, Yip, PK, et al. Increased expression of colony-stimulating factor-1 in mouse spinal cord with experimental autoimmune encephalomyelitis correlates with microglial activation and neuronal loss. Glia. 2018;66:210825. DOI 10.1002/glia.23464.CrossRefGoogle ScholarPubMed
Polgar, E, Hughes, DI, Riddell, JS, Maxwell, DJ, Puskár, Z, Todd, AJ. Selective loss of spinal GABAergic or glycinergic neurons is not necessary for development of thermal hyperalgesia in the chronic constriction injury model of neuropathic pain. Pain. 2003;104:22939. DOI 10.1016/s0304-3959(03)00011-3.CrossRefGoogle Scholar
Polgar, E, Gray, S, Riddell, JS, Todd, AJ. Lack of evidence for significant neuronal loss in laminae I-III of the spinal dorsal horn of the rat in the chronic constriction injury model. Pain. 2004;111:14450. DOI 10.1016/j.pain.2004.06.011.CrossRefGoogle ScholarPubMed
DeLeo, JA, Colburn, RW, Rickman, AJ. Cytokine and growth factor immunohistochemical spinal profiles in two animal models of mononeuropathy. Brain Res. 1997;759:507. DOI 10.1016/s0006-8993(97)00209-6.CrossRefGoogle ScholarPubMed
Brewer, CL, Bacceic, ML. The development of pain circuits and unique effects of neonatal injury. J Neural Transm. 2020;27:46779. DOI 10.1007/s00702-019-02059-z.Google Scholar
Gaudet, AD, Fonken, LK, Ayala, MT, Maier, SF, Watkins, LR. Aging and miR-155 in mice influence survival and neuropathic pain after spinal cord injury. Brain Behav Immun. 2021;97:36570. DOI 10.1016/j.bbi.2021.07.003.CrossRefGoogle ScholarPubMed
Vollert, J, Maier, C, Attal, N, et al. Stratifying patients with peripheral neuropathic pain based on sensory profiles: algorithm and sample size recommendations. Pain. 2017;158:144655. DOI 10.1097/j.pain.0000000000000935.CrossRefGoogle ScholarPubMed
Serra, J, Bostock, H, Sola, R, et al. Microneurographic identification of spontaneous activity in C-nociceptors in neuropathic pain states in humans and rats. Pain. 2012;153:4255. DOI 10.1016/j.pain.2011.08.015.CrossRefGoogle ScholarPubMed
Middleton, SJ, Barry, AM, Comini, M, et al. Studying human nociceptors: from fundamentals to clinic. Brain. 2021;144:131236. DOI 10.1093/brain/awab048.CrossRefGoogle ScholarPubMed
Mogil, JS. Animal models of pain: progress and challenges. Nat Rev Neurosci. 2009;10:28394. DOI 10.1038/nrn2606.CrossRefGoogle ScholarPubMed
Alsaloum, M, Waxman, SG. iPSCs and DRGs: stepping stones to new pain therapies. Trends Mol Med. 2022;28:11022. DOI 10.1016/j.molmed.2021.11.005.CrossRefGoogle ScholarPubMed
Renthal, W, Chamessian, A, Curatolo, M, et al. Human cells and networks of pain: transforming pain target identification and therapeutic development. Neuron. 2021;109:14269. DOI 10.1016/j.neuron.2021.04.005.CrossRefGoogle ScholarPubMed
Bouali-Benazzouz, R, Landry, M, Benazzouz, A, Fossat, P. Neuropathic pain modeling: focus on synaptic and ion channel mechanisms. Prog Neurobiol. 2021;201:102031. DOI 10.1016/jpneurobio.2021.102030.CrossRefGoogle ScholarPubMed