Next Article in Journal
T Cell Receptor Excision Circle (TREC) Monitoring after Allogeneic Stem Cell Transplantation; a Predictive Marker for Complications and Clinical Outcome
Next Article in Special Issue
Procyanidin A2 Modulates IL-4-Induced CCL26 Production in Human Alveolar Epithelial Cells
Previous Article in Journal
Sulforaphane Prevents Testicular Damage in Kunming Mice Exposed to Cadmium via Activation of Nrf2/ARE Signaling Pathways
Previous Article in Special Issue
GSPE Inhibits HMGB1 Release, Attenuating Renal IR-Induced Acute Renal Injury and Chronic Renal Fibrosis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Modulatory Mechanism of Nociceptive Neuronal Activity by Dietary Constituent Resveratrol

1
Laboratory of Food and Physiological Sciences, Department of Life and Food Sciences, School of Life and Environmental Sciences, Azabu University, 1-17-71, Fuchinobe, Chuo-ku, Sagamihara, Kanagawa 252-5201, Japan
2
FANCL Health Science Research Center, Research Institute, FANCL corporation, 12-13, Kamishinano, Totsuka-ku, Yokohama, Kanagawa 244-0806, Japan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2016, 17(10), 1702; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms17101702
Submission received: 2 September 2016 / Revised: 4 October 2016 / Accepted: 5 October 2016 / Published: 11 October 2016
(This article belongs to the Special Issue Natural Anti-Inflammatory Agents)

Abstract

:
Changes to somatic sensory pathways caused by peripheral tissue, inflammation or injury can result in behavioral hypersensitivity and pathological pain, such as hyperalgesia. Resveratrol, a plant polyphenol found in red wine and various food products, is known to have several beneficial biological actions. Recent reports indicate that resveratrol can modulate neuronal excitability, including nociceptive sensory transmission. As such, it is possible that this dietary constituent could be a complementary alternative medicine (CAM) candidate, specifically a therapeutic agent. The focus of this review is on the mechanisms underlying the modulatory effects of resveratrol on nociceptive neuronal activity associated with pain relief. In addition, we discuss the contribution of resveratrol to the relief of nociceptive and/or pathological pain and its potential role as a functional food and a CAM.

Graphical Abstract

1. Introduction

Resveratrol (trans-3,4′,5-trihydroxystilbene) is a dietary constituent and a plant polyphenol found in red wine and various food products [1,2]. It has been reported to have various beneficial actions, including anti-oxidative, anti-inflammatory, neuroprotective, anticancer, and cardioprotective effects [2,3,4,5,6]. Complementary and alternative medicines (CAM), such as herbal medicines and acupuncture, have been used to treat persistent clinical chronic pain [7,8,9], and considerable research has focused on the potential effects of diet and dietary supplementation on conditions associated with pain [10,11,12]. Resveratrol could be a CAM candidate for the treatment of pain.
Recent studies reported that resveratrol modulates neuronal excitability of the peripheral and central nervous systems (PNS and CNS, respectively) via various voltage-dependent [13,14,15,16] and ligand-gated [17,18] ion channels, including the sensory information processing system. Because resveratrol decreases the production of prostaglandin E2 (PGE2) by inhibiting the cyclooxygenase (COX)-2 cascade, resveratrol is a potent inhibitor of inducible COX-2 [19,20]. In fact, PGE2 is a well-known proinflammatory mediator and sensitizer of peripheral nociceptors that also acts on the CNS, including somatosensory neurons in the spinal dorsal horn [21,22,23]. Previous reports indicated that resveratrol inhibits inflammation-induced hyperalgesia by suppressing COX-1 and COX-2 activity [20,24,25]. It is well known that the analgesic action of the acidic antipyretic non-steroidal anti-inflammatory drugs (NSAIDs) involves the potent and efficient inhibition of COX-2 [26]. Together, these observations suggest that resveratrol could be used as a potential therapeutic agent for the prevention of inflammatory hyperalgesia.
The focus of this review is on the mechanism by which resveratrol modulates nociceptive neuronal activity, and its association with the relief of nociceptive and inflammatory pain. In addition, we introduce recent data and discuss the potential contribution of resveratrol to the relief of nociceptive and pathological pain, as well as its development as a functional food and CAM.

2. Trigeminal Pain Pathway and Spinal Trigeminal Nucleus

The trigeminal nervous system is known to have unique structures and functions for the processing of orofacial nociception as well as non-noxious sensations. The oral mucosal membrane, tongue, tooth pulp, gum, and temporomandibular joint (TMJ) are innervated by small-diameter Aδ-fibers and unmyelinated C-fibers that process orofacial nociception [27]. Noxious sensory information in the area innervated by the trigeminal nerves is relayed from trigeminal afferents to second-order neurons in the spinal trigeminal nucleus in the brainstem and the upper cervical (C1–C2) spinal cord [27,28]. The spinal trigeminal nucleus, an important relay station in the transmission of orofacial sensory information, is functionally subdivided into three nuclei: oralis, interpolaris, and caudalis [27]. In addition to the C1–C2 dorsal horn, the spinal trigeminal nucleus caudalis (SpVc) is a relay station for trigeminal nociceptive inputs from inflammation and tissue injury [27,28]. The properties of somatic sensory pathways can be altered by chronic pathological conditions, such as tissue inflammation, leading to hyperalgesia and allodynia [29]. Information processing in the spinal trigeminal nucleus or higher centers is altered by changes in the excitability of primary afferent neurons, which is known as peripheral sensitization [30]. Because it has been reported previously that wide dynamic range (WDR) neurons in the SpVc play an important role in hyperalgesia, allodynia, and/or referred pain associated with orofacial pain [31,32,33,34], in the following sections we focus on studies of the SpVc WDR neuronal activity in the trigeminal pain pathway.

3. Potential Role for Resveratrol in Alleviating Nociceptive Pain

3.1. Peripheral Mechanism

There are four general processes involved in nociceptive sensory signaling: (1) transduction of external stimuli from peripheral terminals; (2) action potential generation; (3) action potential propagation along axons; and (4) transmission to central terminals that form the presynaptic elements of the first synapses in sensory pathways in the CNS [28,35]. Resveratrol has been reported to modulate the excitability of neurons in the PNS by activating voltage-dependent and transient receptor potential (TRP) channels [13,14,15,16,36]. In vitro, mechanical stimuli have been shown to induce mechanosensitive currents via mechanosensitive channels, such as TRP ankyrin 1 (TRPA1), triggering mechanotransduction in trigeminal neurons innervating the inner walls of the anterior eye chamber [37]. In addition, TRPA1 modulates mechanotransduction in primary sensory neurons [38]. Potent inhibition of TRPA1 in vitro and in vivo by resveratrol [36] suggests that resveratrol attenuates the generator potential and inhibits action potential firing via the mechanical transduction process. Moreover, resveratrol has been shown to modulate Na+ and K+ currents in dorsal root ganglion (DRG) neurons associated with action potential generation [13,16], and that resveratrol predominantly inhibits Na+ currents in acutely dissociated DRG neurons, indicating inhibition of action potential generation [13].
Recently, the effects of local subcutaneous injection of resveratrol into the receptive field of SpVc WDR neurons on non-noxious and noxious mechanical stimulation-induced excitability of these neurons were investigated in vivo [39]. In that study, the mean firing rate of SpVc WDR neurons in response to both non-noxious and noxious mechanical stimuli was dose-dependently reduced by resveratrol; this effect was reversible and the mean magnitude of inhibition of the SpVc neuron discharge frequency was almost equal between resveratrol and the local anesthetic lidocaine (1%). These observations suggest that resveratrol injection into the peripheral receptive field suppresses SpVc neuron excitability, possibly by inhibiting Na+ channels in the nociceptive nerve terminals of trigeminal ganglion neurons. Thus, it may be that resveratrol inhibits the excitability of peripheral terminals of the trigeminal nerve by modulating both the noxious mechanical stimulation-induced generator potential and the initiation of action potential processes (Figure 1).

3.2. Central Mechanism

In hippocampal slices, resveratrol has been shown to significantly suppress glutamate-induced currents in post-synaptic CA1 pyramidal neurons without having any presynaptic effects [17]. In addition, Gao et al. [17] indicated that N-methyl-d-aspartate (NMDA) receptors were more sensitive to resveratrol than α-amino-3-hydroxy-5-methyl-4-isoxazole proprionic acid (AMPA) receptors. It has also been reported that action potential duration and l-type Ca2+ currents in ventricular myocytes are reduced by resveratrol [14]. Thus, resveratrol may suppress glutamatergic excitatory synaptic transmission of the SpVc by inhibiting post-synaptic glutamate receptors and presynaptic Ca2+ channels (Figure 1).
Recently, the effects of acute intravenous administration of resveratrol to rats on the excitability of nociceptive WDR SpVc neuronal activity in vivo in response to mechanical stimulation were investigated [40]. In that study, the mean SpVc WDR neuronal firing rate in response to both non-noxious and noxious mechanical stimuli was dose-dependently inhibited by resveratrol in a reversible manner, and the relative magnitude of inhibition by resveratrol of the SpVc WDR neuronal discharge frequency was significantly greater for noxious than non-noxious stimuli. These findings suggest that, in vivo, trigeminal nociceptive transmission in the SpVc is suppressed by acute intravenous resveratrol at the level of secondary neurons, in addition to primary neurons.
The dose-dependent antinociceptive effects of systemic resveratrol appear to be mediated via an opioidergic mechanism, because naloxone pretreatment of rats completely blocked the analgesic effect of resveratrol [41]. Evoked inhibitory GABAergic pre- and post-synaptic potentials in the periaqueductal gray (PAG) are partially inhibited by opiates acting via μ-opioid receptors [42]. Neuronal activity in the PAG is increased after blockade of μ-opioid receptors as a result of GABAergic disinhibition, resulting in the subsequent activation of serotonergic (5-hydroxytrypytamine (5-HT)) neurons in the nucleus raphe magnus, known as the PAG–nucleus raphe magnus–trigeminal pathway [43,44]. Resveratrol has been shown to facilitate 5-HT3 receptor-mediated ion currents [18], and nociceptive stimulation-evoked SpVc/C1 neuron activity is suppressed by conditioning peripheral nerve stimulation via 5-HT3 receptor-mediated GABAergic inhibition [45,46]. Together, these observations suggest that resveratrol suppresses excitatory synaptic transmission of the SpVc via 5-HT3 receptor-mediated GABAergic inhibition and/or endogenous opioidergic mechanisms. However, further studies are needed to elucidate the precise mechanisms involved.

4. Potential Role for Resveratrol in Alleviating Inflammatory Pain

It has been reported that peripheral tissue injury or inflammation of the innervating trigeminal nerve can alter the properties of trigeminal somatic sensory pathways, causing behavioral hypersensitivity and resulting in increased responses to pain caused by noxious stimuli (e.g., hyperalgesia) [29].
Because it has been shown that resveratrol inhibits COX-1 and COX-2 activity [20,24,25], we recently tested the hypothesis that chronic administration of resveratrol would attenuate inflammation-induced hyperexcitability of trigeminal nociceptive neuronal activity associated with hyperalgesia in behavioral and electrophysiological experiments [47]. In that study, the threshold of escape from mechanical stimulation applied to the orofacial area in inflamed rats was significantly lower than in naïve rats, and the lowered mechanical threshold in inflamed rats was returned to control levels following chronic administration of resveratrol. In addition, after resveratrol administration, the mean discharge frequency of SpVc WDR neurons in inflamed rats was significantly decreased in response to both non-noxious and noxious mechanical stimuli, and significant decreases were observed in the inflammation-induced increased spontaneous discharge of SpVc WDR neurons and the frequency and occurrence of noxious pinch-evoked after discharge [47]. Finally, resveratrol administration restored the expanded receptive field of inflamed rats to control levels [47]. These results suggest that chronic administration of resveratrol attenuates inflammation-induced mechanical inflammatory hyperalgesia and that this effect is due primarily to the suppression of the hyperexcitability of SpVc WDR neurons via inhibition of both peripheral and central COX-2 cascade signaling pathways. These findings support the idea that resveratrol could be used as a potential therapeutic agent, or CAM, for the prevention of trigeminal inflammatory hyperalgesia.
It is known that PGE2 appears to facilitate the activation of TRP vanilloid 1 (TRPV1) and tetrodotoxin (TTX)-resistant (TTX-R) Na+ channels [13,29,48], and it has been reported that resveratrol inhibits both TTX-sensitive (TTX-S) and TTX-R Na+ currents in acutely dissociated DRG neurons. It appears that TTX-R Na+ channels (e.g., Nav1.8 and Nav1.9) are selectively expressed in small- and medium-sized DRG neurons [49]. These small DRG neurons are somata that give rise to thinly and unmyelinated C- and Aδ-fibers, which primarily conduct nociceptive stimuli. These Na+ channels can be modulated by activation of adenylate cyclase and increases in cAMP, possibly leading to protein kinase A-dependent phosphorylation of the Na+ channels. In this way, PGE2 produced during an inflammatory response may significantly increase the excitability of nociceptive fibers (peripheral sensitization). Because it has been reported that the increased excitability of small-diameter trigeminal ganglion neurons seen after PGE2 application involves increases in TTX-R Na+ currents [50], it can be assumed that resveratrol inhibits the excitability of small-diameter trigeminal ganglion neurons by suppressing TTX-R Na+ currents induced by the increased production of PGE2. It seems reasonable to speculate that at least part of the peripheral antinociceptive action of resveratrol arises as a result of the prevention of peripheral sensitization, as is the case for antipyretic analgesics (Figure 1).
Conversely, PGE2 can also act in the CNS, namely in the spinal dorsal horn and SpVc neurons, to produce hyperalgesia [21]. Inflammation-induced increases in COX-2 mRNA and protein have been demonstrated in the spinal cord [23,51], where COX-1 and COX-2 are expressed constitutively. Recent evidence indicates that a major stimulus for the induction of COX-2 is the proinflammatory cytokine interleukin-1β, which is found in the periphery as well as in the CNS and is produced in response to inflammation [23,52]. Two possible molecular mechanisms have been proposed to account for PGE2-induced hyperalgesia via actions in the CNS: (1) PGE2 reduces inhibitory glycinergic neurotransmission via a post-synaptic mechanism [53]; and (2) direct depolarization of deep dorsal horn neurons by higher concentrations of PGE2 [22]. Therefore, it is most likely that systemic administration of resveratrol has central antinociceptive effects by suppressing PGE2-induced reductions in inhibitory glycinergic neurotransmission and enhanced depolarization of SpVc neurons through a post-synaptic mechanism. However, further studies are needed to confirm this hypothesis.

5. Functional Significance of Pain Relief by Resveratrol and Future Directions

A widely accepted trigeminal chronic pain model is the complete Freund’s adjuvant (CFA) inflamed rat model [54,55]. Changes in neuronal properties resulting from tissue injury and inflammation of the area innervating the orofacial area can lead to pathological pain, including hyperalgesia and allodynia [33,54]. In a previous study, we reported that TMJ inflammation-induced hyperexcitability of SpVc WDR neurons innervating the facial skin, contributes to ectopic mechanical allodynia of this area [32,33]. Moreover, resveratrol restored the increased mean spontaneous discharge frequency of SpVc WDR neurons in inflamed rats to control levels [47]. Burnstein et al. [56] reported that an ongoing headache (spontaneous pain) is caused by ongoing activity in the SpVc. The origin of this ongoing activity in the central neurons that relay sensory information is of considerable clinical interest, because it has been suggested that it determines the level of post-traumatic injury and chronic pain [57]. More recently, it was shown that the ongoing activity of WDR neurons in the SpVc is driven from the periphery, because microinjection of lidocaine into trigeminal ganglia significantly decreases ongoing activity [58]. Because chronic administration of resveratrol attenuates the spontaneous activity in inflamed rats, these observations together suggest that resveratrol attenuates the spontaneous discharge activity (due to peripheral and/or trigeminal ganglion sensitization and probably contributing to spontaneous pain) of SpVc WDR neurons innervating the facial skin [59].
Although a previous study indicated that dietary grape seed polyphenol extract inhibited TMJ inflammation-induced pain [60], little is known regarding the mechanism by which polyphenols suppress nociceptive neuronal activity. More recently, we reported that in the absence of inflammatory or neuropathic pain, acute intravenous administration of resveratrol suppresses trigeminal sensory transmission, including nociception [40], and so resveratrol may contribute to the suite of CAMs as a therapeutic agent, for the treatment of trigeminal nociceptive pain [7]. In a previous study, under in vivo conditions, systemic administration of resveratrol attenuated inflammation-induced hyperexcitability of trigeminal SpVc neurons associated with hyperalgesia in rats [47]. Recently, there has been increased interest in the use of CAM for the treatment of persistent chronic pain [9,61]. Patients frequently turn to CAM for pain control when other medical treatments are ineffective [8,9]. Recent studies have also focused on the potential effects of diet and dietary supplementation on conditions associated with pain [10,11,62].
Because surgical incisions cause acute pain, and surgery is a potential cause of chronic pain [63,64], it is possible that resveratrol could effectively reduce clinical pain, including postoperative pain [65,66]. In patients with trigeminal neuralgia, local and intravenous administration of lidocaine has been reported to effectively attenuate pain intensity, including allodynia and hyperalgesia [67,68,69]. The results of the different studies into resveratrol contribute to the development of analgesic drugs with fewer side effects for the treatment of pathological pain, including orofacial pain. In particular, the findings from in vivo studies support the idea that resveratrol is a potential therapeutic agent that could be used as an alternative to alleviate nociceptive pain and prevent trigeminal inflammatory hyperalgesia.

6. Concluding Remarks

Recent studies provide evidence that: (1) local resveratrol injection into the peripheral receptive field suppresses the SpVc neuron excitability, possibly by inhibiting generator and action potentials in the nociceptive nerve terminals of trigeminal ganglion neurons; (2) trigeminal sensory transmission, including nociception, is suppressed by acute intravenous resveratrol; and (3) chronic administration of resveratrol attenuates inflammation-induced mechanical hyperalgesia, and this effect is due primarily to the suppression of hyperexcitability of SpVc WDR neurons via inhibition of peripheral and central COX cascade signaling pathways. Together, these findings support the idea that resveratrol may be a potential therapeutic CAM for the alleviation of nociceptive pain and prevention of trigeminal inflammatory hyperalgesia. The pain relief afforded by resveratrol appears to involve modulation of nociceptive neuronal activity.

Author Contributions

Mamoru Takeda conceived and wrote the manuscript. Shiori Takehana and Kenta Sekiguchi performed the experiments reported in the authors’ previously published papers. Yoshiko Kubota helped with the experiments and helped finalize the manuscript. Yoshihito Shimazu interpreted the data and helped finalize the manuscript. All authors read and approved the final manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Frémont, L. Biological effects of resveratrol. Life Sci. 2000, 66, 663–673. [Google Scholar] [CrossRef]
  2. Pervaiz, S. Resveratrol: From grapevines to mammalian biology. FASEB J. 2003, 17, 1975–1985. [Google Scholar] [CrossRef] [PubMed]
  3. Leiro, J.; Arranz, J.; Fraiz, N.; Sanmartin, M.; Quezada, E.; Orallo, F. Effect of cis-resveratrol on genes involved in nuclear factor kappa B signaling. Int. Immunopharmacol. 2005, 5, 393–406. [Google Scholar] [CrossRef] [PubMed]
  4. Bermúdez-Ocaña, D.Y.; Ambriz-Tututi, M.; Pérez-Severiano, F.; Granados-Soto, V. Pharmacological evidence for the participation of NO–cyclic GMP–PKG–K+ channel pathway in the antiallodynic action of resveratrol. Pharmacol. Biochem. Behav. 2006, 84, 535–542. [Google Scholar] [CrossRef] [PubMed]
  5. Pérez-Severiano, F.; Bermúdez-Ocaña, D.Y.; López-Sánchez, P.; Ríos, C.; Granados-Soto, V. Spinal nerve ligation reduces nitric oxide synthase activity and expression: Effect of resveratrol. Pharmacol. Biochem. Behav. 2008, 90, 742–747. [Google Scholar] [CrossRef] [PubMed]
  6. Russo, G.L. Ins and outs of dietary phytochemicals in cancer chemoprevention. Biochem. Pharmacol. 2007, 74, 533–544. [Google Scholar] [CrossRef] [PubMed]
  7. Rao, J.K.; Mihaliak, K.; Kroenke, K.; Bradley, J.; Tierney, W.M.; Weinberger, M. Use of complementary therapies for arthritis among patients of rheumatologists. Ann. Intern. Med. 1999, 131, 409–416. [Google Scholar] [CrossRef] [PubMed]
  8. Konvicka, J.J.; Meyer, T.A.; McDavid, A.J.; Roberson, C.R. Complementary/alternative medicine use among chronic pain clinic patients. J. PeriAnesthesia Nurs. 2008, 23, 17–23. [Google Scholar] [CrossRef] [PubMed]
  9. Rosenberg, E.I.; Genao, I.; Chen, I.; Mechaber, A.J.; Wood, J.A.; Faselis, C.J.; Kurz, J.; Menon, M.; O’Rorke, J.; Panda, M. Complementary and alternative medicine use by primary care patients with chronic pain. Pain Med. 2008, 9, 1065–1072. [Google Scholar] [CrossRef] [PubMed]
  10. Shir, Y.; Raja, S.N.; Weissman, C.S.; Campbell, J.N.; Seltzer, Z.E. Consumption of soy diet before nerve injury preempts the development of neuropathic pain in rats. J. Am. Soc. Anesthesiol. 2001, 95, 1238–1244. [Google Scholar] [CrossRef]
  11. Ernst, E. Complementary medicine. Curr. Opin. Rheumatol. 2003, 15, 151–155. [Google Scholar] [CrossRef] [PubMed]
  12. Tall, J.M.; Raja, S.N. Dietary constituents as novel therapies for pain. Clin. J. Pain 2004, 20, 19–26. [Google Scholar] [CrossRef] [PubMed]
  13. Im Kim, H.; Kim, T.H.; Song, J.-H. Resveratrol inhibits Na+ currents in rat dorsal root ganglion neurons. Brain Res. 2005, 1045, 134–141. [Google Scholar] [CrossRef] [PubMed]
  14. Liew, R.; Stagg, M.A.; MacLeod, K.T.; Collins, P. The red wine polyphenol, resveratrol, exerts acute direct actions on guinea-pig ventricular myocytes. Eur. J. Pharmacol. 2005, 519, 1–8. [Google Scholar] [CrossRef] [PubMed]
  15. Gao, Z.-B.; Hu, G.-Y. Trans-resveratrol, a red wine ingredient, inhibits voltage-activated potassium currents in rat hippocampal neurons. Brain Res. 2005, 1056, 68–75. [Google Scholar] [CrossRef] [PubMed]
  16. Granados-Soto, V.; Argüelles, C.; Ortiz, M. The peripheral antinociceptive effect of resveratrol is associated with activation of potassium channels. Neuropharmacology 2002, 43, 917–923. [Google Scholar] [CrossRef]
  17. Gao, Z.-B.; Chen, X.-Q.; Hu, G.-Y. Inhibition of excitatory synaptic transmission by trans-resveratrol in rat hippocampus. Brain Res. 2006, 1111, 41–47. [Google Scholar] [CrossRef] [PubMed]
  18. Lee, B.-H.; Hwang, S.-H.; Choi, S.-H.; Shin, T.-J.; Kang, J.; Lee, S.-M.; Nah, S.-Y. Resveratrol enhances 5-hydroxytryptamine type 3A receptor-mediated ion currents: The role of arginine 222 residue in pre-transmembrane domain I. Biol. Pharm. Bull. 2011, 34, 523–527. [Google Scholar] [CrossRef] [PubMed]
  19. Subbaramaiah, K.; Chung, W.J.; Michaluart, P.; Telang, N.; Tanabe, T.; Inoue, H.; Jang, M.; Pezzuto, J.M.; Dannenberg, A.J. Resveratrol inhibits cyclooxygenase-2 transcription and activity in phorbol ester-treated human mammary epithelial cells. J. Biol. Chem. 1998, 273, 21875–21882. [Google Scholar] [CrossRef] [PubMed]
  20. Pham-Marcou, T.A.; Beloeil, H.; Sun, X.; Gentili, M.; Yaici, D.; Benoit, G.; Benhamou, D.; Mazoit, J.-X. Antinociceptive effect of resveratrol in carrageenan-evoked hyperalgesia in rats: Prolonged effect related to COX-2 expression impairment. Pain 2008, 140, 274–283. [Google Scholar] [CrossRef] [PubMed]
  21. Neugebauer, V.; Geisslinger, G.; Rümenapp, P.; Weiretter, F.; Szelenyi, I.; Brune, K.; Schaible, H.-G. Antinociceptive effects of R(−)- and S(+)-flurbiprofen on rat spinal dorsal horn neurons rendered hyperexcitable by an acute knee joint inflammation. J. Pharm. Exp. Ther. 1995, 275, 618–628. [Google Scholar]
  22. Baba, H.; Kohno, T.; Moore, K.A.; Woolf, C.J. Direct activation of rat spinal dorsal horn neurons by prostaglandin E2. J. Neurosci. 2001, 21, 1750–1756. [Google Scholar] [PubMed]
  23. Samad, T.A.; Moore, K.A.; Sapirstein, A.; Billet, S.; Allchorne, A.; Poole, S.; Bonventre, J.V.; Woolf, C.J. Interleukin-1β-mediated induction of Cox-2 in the CNS contributes to inflammatory pain hypersensitivity. Nature 2001, 410, 471–475. [Google Scholar] [CrossRef] [PubMed]
  24. Torres-López, J.E.; Ortiz, M.I.; Castaneda-Hernández, G.; Alonso-López, R.; Asomoza-Espinosa, R.; Granados-Soto, V. Comparison of the antinociceptive effect of celecoxib, diclofenac and resveratrol in the formalin test. Life Sci. 2002, 70, 1669–1676. [Google Scholar] [CrossRef]
  25. Gentilli, M.; Mazoit, J.X.; Bouaziz, H.; Fletcher, D.; Casper, R.F.; Benhamou, D.; Savouret, J.-F. Resveratrol decreases hyperalgesia induced by carrageenan in the rat hind paw. Life Sci. 2001, 68, 1317–1321. [Google Scholar] [CrossRef]
  26. Brune, K.; Zeilhofer, H. Antipyretic analgesics: Basic aspects. In Textbook of Pain; Elsevier: Amsterdam, The Netherlands, 2006; pp. 459–469. [Google Scholar]
  27. Sessle, B.J. Acute and chronic craniofacial pain: Brainstem mechanisms of nociceptive transmission and neuroplasticity, and their clinical correlates. Crit. Rev. Oral Biol. Med. 2000, 11, 57–91. [Google Scholar] [CrossRef] [PubMed]
  28. Takeda, M.; Matsumoto, S.; Sessle, B.J.; Shinoda, M.; Iwata, K. Peripheral and central mechanisms of trigeminal neuropathic and inflammatory pain. J. Oral Biosci. 2011, 53, 318–329. [Google Scholar] [CrossRef]
  29. Scholz, J.; Woolf, C.J. Can we conquer pain? Nat. Neurosci. 2002, 5, 1062–1067. [Google Scholar] [CrossRef] [PubMed]
  30. Millan, M.J. The induction of pain: An integrative review. Prog. Neurobiol. 1999, 57, 1–164. [Google Scholar] [CrossRef]
  31. Takeda, M.; Tanimoto, T.; Matsumoto, S. Change in mechanical receptive field properties induced by GABAA receptor activation in the trigeminal spinal nucleus caudalis neurons in rats. Exp. Brain Res. 2000, 134, 409–416. [Google Scholar] [CrossRef] [PubMed]
  32. Takeda, M.; Tanimoto, T.; Ito, M.; Nasu, M.; Matsumoto, S. Role of capsaicin-sensitive primary afferent inputs from the masseter muscle in the C1 spinal neurons responding to tooth-pulp stimulation in rats. Exp. Brain Res. 2005, 160, 107–117. [Google Scholar] [CrossRef] [PubMed]
  33. Takeda, M.; Takahashi, M.; Matsumoto, S. Suppression of neurokinin-1 receptor in trigeminal ganglia attenuates central sensitization following inflammation. J. Peripher. Nerv. Syst. 2012, 17, 169–181. [Google Scholar] [CrossRef] [PubMed]
  34. Nishikawa, T.; Takeda, M.; Tanimoto, T.; Matsumoto, S. Convergence of nociceptive information from temporomandibular joint and tooth pulp afferents on C1 spinal neurons in the rat. Life Sci. 2004, 75, 1465–1478. [Google Scholar] [CrossRef] [PubMed]
  35. Harriott, A.M.; Gold, M.S. Contribution of primary afferent channels to neuropathic pain. Curr. Pain Headache Rep. 2009, 13, 197–207. [Google Scholar] [CrossRef] [PubMed]
  36. Yu, L.; Wang, S.; Kogure, Y.; Yamamoto, S.; Noguchi, K.; Dai, Y. Modulation of TRP channels by resveratrol and other stilbenoids. Mol. Pain 2013, 9, 1. [Google Scholar] [CrossRef] [PubMed]
  37. Meng, Q.; Fang, P.; Hu, Z.; Ling, Y.; Liu, H. Mechanotransduction of trigeminal ganglion neurons innervating inner walls of rat anterior eye chambers. Am. J. Physiol.–Cell Physiol. 2015, 309, C1–C10. [Google Scholar] [CrossRef] [PubMed]
  38. Kwan, K.Y.; Glazer, J.M.; Corey, D.P.; Rice, F.L.; Stucky, C.L. TRPA1 modulates mechanotransduction in cutaneous sensory neurons. J. Neurosci. 2009, 29, 4808–4819. [Google Scholar] [CrossRef] [PubMed]
  39. Shimazu, Y.; Shibuya, E.; Takehana, S.; Sekiguchi, K.; Oshima, K.; Kamata, H.; Karibe, H.; Takeda, M. Local administration of resveratrol inhibits excitability of nociceptive wide-dynamic range neurons in rat trigeminal spinal nucleus caudalis. Brain Res. Bull. 2016, 124, 262–268. [Google Scholar] [CrossRef] [PubMed]
  40. Takehana, S.; Sekiguchi, K.; Inoue, M.; Kubota, Y.; Ito, Y.; Yui, K.; Shimazu, Y.; Takeda, M. Systemic administration of resveratrol suppress the nociceptive neuronal activity of spinal trigeminal nucleus caudalis in rats. Brain Res. Bull. 2016, 120, 117–122. [Google Scholar] [CrossRef] [PubMed]
  41. Gupta, Y.; Sharma, M.; Briyal, S. Antinociceptive effect of trans-resveratrol in rats: Involvement of an opioidergic mechanism. Methods Find. Exp. Clin. Pharmacol. 2004, 26, 667–672. [Google Scholar] [CrossRef] [PubMed]
  42. Chieng, B.; Christie, M. Hyperpolarization by opioids acting on μ-receptors of a sub-population of rat periaqueductal gray neurones in vitro. Br. J. Pharmacol. 1994, 113, 121–128. [Google Scholar] [CrossRef] [PubMed]
  43. Gebahrts, G.F.; Randich, A. Brain stem modulation of nociception. In Brain Stem Mechanisms of Behavior; Klemm, W.R., Vertes, R.P., Eds.; Wiley-Intersciences: New York, NY, USA, 1990; pp. 315–352. [Google Scholar]
  44. Takeda, M.; Tanimoto, T.; Nishikawa, T.; Ikeda, M.; Yoshida, S.; Ito, M.; Matsumoto, S. Volume expansion suppresses the tooth-pulp evoked jaw-opening reflex related activity of trigeminal neurons in rats. Brain Res. Bull. 2002, 58, 83–89. [Google Scholar] [CrossRef]
  45. Tanimoto, T.; Takeda, M.; Nishikawa, T.; Matsumoto, S. The role of 5-HT3 receptors in the vagal afferent activation-induced inhibition of C1 spinal neurons projected from tooth-pulp in the rat. J. Pharmacol. Exp. Ther. 2004. [Google Scholar] [CrossRef] [PubMed]
  46. Oshima, K.; Takeda, M.; Tanimoto, T.; Katsuumi, I.; Matsumoto, S. Tooth-pulp-evoked rostral spinal trigeminal nucleus neuron activity is inhibited by conditioning sciatic nerve stimulation in the rat: Possible role of 5-HT3 receptor mediated GABAergic inhibition. Brain Res. Bull. 2005, 65, 31–40. [Google Scholar] [CrossRef] [PubMed]
  47. Sekiguchi, K.; Takehana, S.; Shibuya, E.; Matsuzawa, N.; Hidaka, S.; Kanai, Y.; Inoue, M.; Kubota, Y.; Shimazu, Y.; Takeda, M. Resveratrol attenuates inflammation-induced hyperexcitability of trigeminal spinal nucleus caudalis neurons associated with hyperalgesia in rats. Mol. Pain 2016, 12. [Google Scholar] [CrossRef] [PubMed]
  48. Caterina, M.J.; Schumacher, M.A.; Tominaga, M.; Rosen, T.A.; Levine, J.D.; Julius, D. The capsaicin receptor: A heat-activated ion channel in the pain pathway. Nature 1997, 389, 816–824. [Google Scholar] [PubMed]
  49. Akopian, A.N.; Sivilotti, L.; Wood, J.N. A tetrodotoxin-resistant voltage-gated sodium channel expressed by sensory neurons. Nature 1996, 379, 257–262. [Google Scholar] [CrossRef] [PubMed]
  50. Kadoi, J.; Takeda, M.; Matsumoto, S. Prostaglandin E2 potentiates the excitability of small diameter trigeminal root ganglion neurons projecting onto the superficial layer of the cervical dorsal horn in rats. Exp. Brain Res. 2007, 176, 227–236. [Google Scholar] [CrossRef] [PubMed]
  51. Beiche, F.; Scheuerer, S.; Brune, K.; Geisslinger, G.; Goppelt-Struebe, M. Up-regulation of cyclooxygenase-2 mRNA in the rat spinal cord following peripheral inflammation. FEBS Lett. 1996, 390, 165–169. [Google Scholar] [CrossRef]
  52. Ek, T.; Jarfelt, M.; Mellander, L.; Abrahamsson, J. Proinflammatory cytokines mediate the systemic inflammatory response associated with high-dose cytarabine treatment in children. Med. Pediatr. Oncol. 2001, 37, 459–464. [Google Scholar] [CrossRef] [PubMed]
  53. Ahmadi, S.; Lippross, S.; Neububer, W.L.; Zeilhofer, H.U. PGE2 selectively blocks inhibitory glycinergic neurotransmission onto rat superficial dorsal horn neurons. Nat. Neurosci. 2002, 5, 34–40. [Google Scholar]
  54. Iwata, K.; Tashiro, A.; Tsuboi, Y.; Imai, T.; Sumino, R.; Morimoto, T.; Dubner, R.; Ren, K. Medullary dorsal horn neuronal activity in rats with persistent temporomandibular joint and perioral inflammation. J. Neurophysiol. 1999, 82, 1244–1253. [Google Scholar] [PubMed]
  55. Imbe, H.; Iwata, K.; Zou, S.; Dubner, R.; Ren, K. Orofacial deep and cutaneous tissue inflammation and trigeminal neuronal activation. Cells Tissues Organs 2001, 169, 238–247. [Google Scholar] [CrossRef] [PubMed]
  56. Burstein, R.; Curter, M.F.; Yarnitsky, D. The development of cutaneous allodynia during a migraine attack: Clinical evidence for the sequential of spinal and supraspinal nociceptive neurons in maigraine. Brain 2000, 123, 1703–1709. [Google Scholar]
  57. Sorkin, L.S.; Wallace, M.S. Acute pain mechanisms. Surg. Clin. N. Am. 1999, 79, 213–229. [Google Scholar] [CrossRef]
  58. Roch, M.; Messlinger, K.; Kulchitsky, V.; Tichonovich, O.; Azev, O.; Koulchitsky, S. Ongoing activity in trigeminal wide-dynamic range neurons is driven from the periphery. Neuroscience 2007, 150, 681–691. [Google Scholar] [CrossRef] [PubMed]
  59. Takeda, M.; Takahashi, M.; Matsumoto, S. Contribution of the activation of satellite glia in sensory ganglia to pathological pain. Neurosci. Biobehav. Rev. 2009, 33, 784–792. [Google Scholar] [CrossRef] [PubMed]
  60. Cady, R.J.; Hirst, J.J.; Durham, P.L. Dietary grape seed polyphenols repress neuron and glia activation in trigeminal ganglion and trigeminal nucleus caudalis. Mol. Pain 2010, 6, 1. [Google Scholar] [CrossRef] [PubMed]
  61. Kessler, R.C.; Davis, R.B.; Foster, D.F.; van Rompay, M.I.; Walters, E.E.; Wilkey, S.A.; Kaptchuk, T.J.; Eisenberg, D.M. Long-term trends in the use of complementary and alternative medical therapies in the United States. Ann. Intern. Med. 2001, 135, 262–268. [Google Scholar] [CrossRef] [PubMed]
  62. Rivat, C.; Richebé, P.; Laboureyras, E.; Laulin, J.-P.; Havouis, R.; Noble, F.; Moulinoux, J.-P.; Simonnet, G. Polyamine deficient diet to relieve pain hypersensitivity. PAIN® 2008, 137, 125–137. [Google Scholar] [CrossRef] [PubMed]
  63. Perkins, F.M.; Kehlet, H. Chronic pain as an outcome of surgery: A review of predictive factors. J. Am. Soc. Anesthesiol. 2000, 93, 1123–1133. [Google Scholar] [CrossRef]
  64. Kehlet, H.; Jensen, T.S.; Woolf, C.J. Persistent postsurgical pain: Risk factors and prevention. Lancet 2006, 367, 1618–1625. [Google Scholar] [CrossRef]
  65. Locher-Claus, M.T.; Erickson, T.E.; Law, A.S.; Johnson, W.T.; Gebhart, G. Effects of pre-emptive morphine, ibuprofen or local anesthetic on Fos expression in the spinal trigeminal nucleus following tooth pulp exposure in the rat. J. Endod. 2005, 31, 578–583. [Google Scholar] [CrossRef] [PubMed]
  66. Tillu, D.V.; Melemedjian, O.K.; Asiedu, M.N.; Qu, N.; de Felice, M.; Dussor, G.; Price, T.J. Resveratrol engages AMPK to attenuate ERK and mTOR signaling in sensory neurons and inhibits incision-induced acute and chronic pain. Mol. Pain 2012, 8, 1. [Google Scholar] [CrossRef] [PubMed]
  67. Kanai, A.; Suzuki, A.; Kobayashi, M.; Hoka, S. Intranasal lidocaine 8% spray for second-division trigeminal neuralgia. Br. J. Anaesth. 2006, 97, 559–563. [Google Scholar] [CrossRef] [PubMed]
  68. Han, K.R.; Kim, C.; Chae, Y.; Kim, D.W. Efficacy and safety of high concentration lidocaine for trigeminal nerve block in patients with trigeminal neuralgia. Int. J. Clin. Pract. 2008, 62, 248–254. [Google Scholar] [CrossRef] [PubMed]
  69. Stavropoulou, E.; Argyra, E.; Zis, P.; Vadalouca, A.; Siafaka, I. The effect of intravenous lidocaine on trigeminal neuralgia: A randomized double blind placebo controlled trial. ISRN Pain 2014, 2014. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic drawing of the possible mechanism underlying the effects of resveratrol in relieving (a) nociceptive and(b) inflammatory pain. (a) nociceptive pain. When noxious mechanical stimulation is applied to the skin, mechanosensitive ion channels (e.g., transient receptor potential ankyrin 1 (TRPA1) channels) open, activating the generator potential (depolarization). This depolarization further opens voltage-dependent sodium and potassium channels, generating action potentials. Action potentials are discharged through primary afferent fibers (Aδ- and C-fibers) to the central terminal in nociceptive neurons in the trigeminal spinal nucleus caudalis (SpVc). When action potentials are conducted to the central terminal of the SpVc, presynaptic voltage-dependent calcium channels open, leading to the release of neurotransmitters (e.g., glutamate) into the synaptic cleft, which then bind to post-synaptic (glutamate) receptors, activating excitatory post-synaptic potentials (EPSP). If the amplitude of EPSPs is over the action potential threshold, a barrage of action potentials is conducted to higher centers in the pain pathway and pain is perceived. It is possible that resveratrol suppresses both the excitability of peripheral terminals of the trigeminal nerve (by modulating both the mechanical transduction and generation of action potentials) and glutaminergic excitatory synaptic transmission of the SpVc (by inhibiting post-synaptic glutamate receptors and presynaptic Ca2+ channels). R, resveratrol; Nav, voltage-gated sodium channel; Kv, voltage-gated potassium channel; Cav, voltage-gated calcium channel; Glu, glutamate; WDR, wide dynamic range neurons; (b) Inflammatory pain. Following peripheral inflammation and/or nerve injury, inflammatory mediators, such as prostaglandin E2 (PGE2), bind to G-protein-coupled E-type prostanoid (EP) receptors and induce activation of protein kinases A and C (PKA and PKC, respectively) in nociceptive peripheral terminals, leading to phosphorylation of mechanosensitive, sodium and potassium ion channels and receptors. As a result, the activation threshold for transducer channels such as TRPA1 is reduced and the membrane excitability of the peripheral terminals increases, resulting in a high frequency of action potentials being conducted to presynaptic central terminals of the SpVc. This results in the release of a large amount of glutamate into the synaptic cleft, which binds to upregulated post-synaptic glutamate receptors, augmenting EPSPs, causing a barrage of action potentials to be conducted to higher centers of pain pathways and creating a state of heightened sensitivity termed peripheral sensitization. It is possible that chronic administration of resveratrol attenuates inflammation-induced mechanical inflammatory hyperalgesia, with this effect due primarily to suppression of the hyperexcitability of SpVc WDR neurons via inhibition of both peripheral and central cyclooxygenase (COX)-2 cascade signaling pathways. ARA, arachidonic acid. X: Suppression.
Figure 1. Schematic drawing of the possible mechanism underlying the effects of resveratrol in relieving (a) nociceptive and(b) inflammatory pain. (a) nociceptive pain. When noxious mechanical stimulation is applied to the skin, mechanosensitive ion channels (e.g., transient receptor potential ankyrin 1 (TRPA1) channels) open, activating the generator potential (depolarization). This depolarization further opens voltage-dependent sodium and potassium channels, generating action potentials. Action potentials are discharged through primary afferent fibers (Aδ- and C-fibers) to the central terminal in nociceptive neurons in the trigeminal spinal nucleus caudalis (SpVc). When action potentials are conducted to the central terminal of the SpVc, presynaptic voltage-dependent calcium channels open, leading to the release of neurotransmitters (e.g., glutamate) into the synaptic cleft, which then bind to post-synaptic (glutamate) receptors, activating excitatory post-synaptic potentials (EPSP). If the amplitude of EPSPs is over the action potential threshold, a barrage of action potentials is conducted to higher centers in the pain pathway and pain is perceived. It is possible that resveratrol suppresses both the excitability of peripheral terminals of the trigeminal nerve (by modulating both the mechanical transduction and generation of action potentials) and glutaminergic excitatory synaptic transmission of the SpVc (by inhibiting post-synaptic glutamate receptors and presynaptic Ca2+ channels). R, resveratrol; Nav, voltage-gated sodium channel; Kv, voltage-gated potassium channel; Cav, voltage-gated calcium channel; Glu, glutamate; WDR, wide dynamic range neurons; (b) Inflammatory pain. Following peripheral inflammation and/or nerve injury, inflammatory mediators, such as prostaglandin E2 (PGE2), bind to G-protein-coupled E-type prostanoid (EP) receptors and induce activation of protein kinases A and C (PKA and PKC, respectively) in nociceptive peripheral terminals, leading to phosphorylation of mechanosensitive, sodium and potassium ion channels and receptors. As a result, the activation threshold for transducer channels such as TRPA1 is reduced and the membrane excitability of the peripheral terminals increases, resulting in a high frequency of action potentials being conducted to presynaptic central terminals of the SpVc. This results in the release of a large amount of glutamate into the synaptic cleft, which binds to upregulated post-synaptic glutamate receptors, augmenting EPSPs, causing a barrage of action potentials to be conducted to higher centers of pain pathways and creating a state of heightened sensitivity termed peripheral sensitization. It is possible that chronic administration of resveratrol attenuates inflammation-induced mechanical inflammatory hyperalgesia, with this effect due primarily to suppression of the hyperexcitability of SpVc WDR neurons via inhibition of both peripheral and central cyclooxygenase (COX)-2 cascade signaling pathways. ARA, arachidonic acid. X: Suppression.
Ijms 17 01702 g001

Share and Cite

MDPI and ACS Style

Takeda, M.; Takehana, S.; Sekiguchi, K.; Kubota, Y.; Shimazu, Y. Modulatory Mechanism of Nociceptive Neuronal Activity by Dietary Constituent Resveratrol. Int. J. Mol. Sci. 2016, 17, 1702. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms17101702

AMA Style

Takeda M, Takehana S, Sekiguchi K, Kubota Y, Shimazu Y. Modulatory Mechanism of Nociceptive Neuronal Activity by Dietary Constituent Resveratrol. International Journal of Molecular Sciences. 2016; 17(10):1702. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms17101702

Chicago/Turabian Style

Takeda, Mamoru, Shiori Takehana, Kenta Sekiguchi, Yoshiko Kubota, and Yoshihito Shimazu. 2016. "Modulatory Mechanism of Nociceptive Neuronal Activity by Dietary Constituent Resveratrol" International Journal of Molecular Sciences 17, no. 10: 1702. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms17101702

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