Next Article in Journal
COVID-19 and Bell’s Palsy
Previous Article in Journal
The Post-Pandemic Transformation of Art and Architecture Libraries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Entry

Inhibitory Actions of Clinical Analgesics, Analgesic Adjuvants, and Plant-Derived Analgesics on Nerve Action Potential Conduction

Faculty of Medicine, Saga University, 5-1-1 Nabeshima, Saga 849-8501, Japan
Encyclopedia 2022, 2(4), 1902-1934; https://doi.org/10.3390/encyclopedia2040132
Submission received: 22 August 2022 / Revised: 19 October 2022 / Accepted: 23 October 2022 / Published: 30 November 2022
(This article belongs to the Section Medicine & Pharmacology)

Definition

:
The action potential (AP) conduction in nerve fibers plays a crucial role in transmitting nociceptive information from the periphery to the cerebral cortex. Nerve AP conduction inhibition possibly results in analgesia. It is well-known that many analgesics suppress nerve AP conduction and voltage-dependent sodium and potassium channels that are involved in producing APs. The compound action potential (CAP) recorded from a bundle of nerve fibers is a guide for knowing if analgesics affect nerve AP conduction. This entry mentions the inhibitory effects of clinically used analgesics, analgesic adjuvants, and plant-derived analgesics on fast-conducting CAPs and voltage-dependent sodium and potassium channels. The efficacies of their effects were compared among the compounds, and it was revealed that some of the compounds have similar efficacies in suppressing CAPs. It is suggested that analgesics-induced nerve AP conduction inhibition may contribute to at least a part of their analgesic effects.

1. Introduction

Signal of painful stimuli applied to the skin is mainly conveyed by primary-afferent thin myelinated Aδ-fibers and unmyelinated C-fibers to the spinal cord and brain stem; the information is then transmitted to the brain by the conduction of action potentials (APs) in nerve fibers and chemical transmission at neuron-to-neuron junctions [1,2,3,4]. Acute nociceptive pain caused by tissue injury or damage is a physiological mechanism that serves to protect a person against injury, which is usually alleviated by antipyretic analgesics including non-steroidal anti-inflammatory drugs (NSAIDs) and narcotic analgesics such as opioids. On the other hand, chronic pain, which may last for a long time, such as three months or more, or occur repeatedly, is a debilitating disease accompanied by spontaneous pain, etc. and is often resistant to analgesics such as NSAIDs and opioids. Neuropathic pain, which is one type of chronic pain, results from a direct injury given to the peripheral nervous system (PNS) and damage caused in the central nervous system (CNS), and it is characterized by an excessive rise in the excitability of neurons in the vicinity of injured or damaged neuronal tissues [5]. This type of pain is alleviated by using analgesic adjuvants such as local anesthetics, antiepileptics, antidepressants, and α2-adrenoceptor agonists [6,7,8,9,10,11,12,13,14,15]. Although analgesics and analgesic adjuvants generally depress excitatory synaptic transmission [16,17,18], many of their drugs can possibly suppress nerve AP conduction, which in part contributes to their inhibitory effects on pain. Plants and their constituents are used as folk remedies to relieve pain as a drug with few side effects [19,20,21].
AP conduction is produced by the activation of voltage-dependent sodium and potassium channels expressed in nerve fibers. Thus, a stimulus that induces membrane depolarization, applied to a certain point on a nerve fiber, opens a sodium channel, resulting in an influx of sodium ion into the cell according to the concentration and potential gradient across the cell membrane. This leads to AP production in a self-renewing manner, which in turn causes an outward current, i.e., membrane depolarization, to open other sodium channels at the points next to it. Such an AP production is subsided by subsequent sodium channel inactivation and potassium channel opening [22,23].

Isolation Methods for Testing Analgesic Action on Nerve Fibers

The study of AP in mammals is complicated by the need to dissect out individual nerves to isolate them from peripheral stimulation. For this reason, neuroscience relies on nerve extraction from relatively simple animals with conserved AP mechanisms, such as insects, reptiles, squids, or frogs (for example, refer to [23]).
AP current flowing on the surface of a nerve trunk consisting of many fibers can be measured as a compound action potential (CAP) by immersing the nerve in an isolator such as air, oil, or sucrose, and then by putting two electrodes on the nerve. CAPs, which are sensitive to tetrodotoxin (TTX), that blocks voltage-dependent sodium channels, and are fast-conducting (possibly mediated by primary-afferent thick myelinated Aα fibers), can be easily observed in the sciatic nerve trunk isolated from frogs by exposing the nerve trunk to air (known as the air-gap method). A half-peak duration of the CAP was increased by a voltage-dependent delayed rectifier potassium channel inhibitor, tetraethylammonium, without any alteration in its peak amplitude, which indicated that potassium channels are involved in CAP production [24]. Although the frog sciatic nerve exhibits both fast-conducting and slow-conducting (Aδ-fiber and C-fiber mediated) CAPs, the latter CAPs have much smaller peak amplitudes and conduction velocities than the former ones [25].
Fast-conducting CAPs recorded from the frog sciatic nerve were found to be inhibited by antinociceptive drugs in a manner dependent on their concentrations and chemical structures. Among the drugs, there are clinically used antinociceptive drugs including NSAIDs [26], many types of opioids such as tramadol [27,28], many amide- and ester-type local anesthetics [29], antiepileptics [30], antidepressants [31], dexmedetomidine (DEX; (+)-(S)-4-[1-(2,3-dimethylphenyl)-ethyl]-1H-imidazole, which is an α2-adrenoceptor agonist; [32]), and diverse kinds of antinociceptive compounds isolated from plants [33]. This entry will describe the effects of antinociceptive drugs on CAPs evoked in the sciatic nerves of frogs and argue how nerve AP conduction inhibitions produced by drugs differ among them. For comparison, the effects of antinociceptive drugs on peripheral nerve CAPs in mammals and voltage-dependent sodium and potassium channels that are involved in producing APs will also be mentioned, provided that data are available.

2. Actions of Analgesics on Nerve AP Conduction

2.1. NSAIDs

NSAIDs globally downregulate nociception through multiple mechanisms, including: inhibited production of prostaglandins from arachidonic acid by the suppression of the cyclooxygenase enzyme ([34,35]; refer to reviews [36,37,38,39]); activation of several potassium channels ([40,41,42,43,44]; refer to reviews [45,46]); suppression of acid-sensitive ion channels [47] and transient receptor potential (TRP) channels [48,49]; substance P depletion [50]; interaction with the adrenergic nervous system [51]; opioid receptor activation [52,53]; and cannabinoid receptor activation and increase in endocannabinoid level [54]. The idea that antinociception is produced by effects other than the suppression of cyclooxygenase is consistent with the experimental result which showed a difference between the antinociceptive and anti-inflammatory effects of NSAIDs [55].
Diclofenac, which is an acetic acid-derived NSAID, inhibited frog sciatic nerve CAPs in a partially reversible fashion; a half-maximal inhibitory concentration (IC50) value for this activity was 0.94 mM at concentrations ranging from 0.01 to 1 mM. Like diclofenac, another acetic acid-derived NSAID, aceclofenac, which is a carboxymethyl ester of diclofenac, was shown to depress CAPs at concentrations ranging from 0.01 to 1 mM; an IC50 value for this activity was 0.47 mM, which is less than that of diclofenac. Other acetic acid-derived NSAIDs exhibited a similar CAP inhibition, albeit to smaller extents compared with diclofenac and aceclofenac. Indomethacin at 1 mM suppressed the peak amplitude of CAP with an extent of 38%. Acemetacin, in which the -OH group present in indomethacin is replaced with -OCH2COOH, reduced CAP peak amplitudes by 38% at 0.5 mM concentration. Etodolac at 1 mM suppressed the peak amplitude of CAP with an extent of only 15%, and sulindac and felbinac at 1 mM did not affect CAP peak amplitudes [26].
Frog sciatic nerve CAP peak amplitudes were concentration-dependently reduced by fenamic acid-derived NSAIDs such as tolfenamic acid, meclofenamic acid, mefenamic acid, and flufenamic acid, the chemical structures of which resemble those of diclofenac and aceclofenac. Tolfenamic acid reduced CAP peak amplitudes with an IC50 value of 0.29 mM at concentrations ranging from 0.01 to 0.2 mM. Meclofenamic acid, which has the -Cl group present in different numbers and positions, attached to the benzene ring of tolfenamic acid, had an IC50 value of 0.19 mM when examined at concentrations ranging from 0.01 to 0.5 mM. Furthermore, mefenamic acid, in which the -Cl group attached to the benzene ring of tolfenamic acid is substituted with the -CH3 group, reduced the peak amplitude of CAP at concentrations ranging from 0.01 to 0.2 mM (by 16% at 0.2 mM). Flufenamic acid, in which one of two -CH3 groups attached to the benzene ring of mefenamic acid is absent and the other is substituted with the -CF3 group, exhibited an IC50 value of 0.22 mM, which is similar to the values exhibited by tolfenamic acid and meclofenamic acid [26]. With respect to other types of NSAIDs, frog sciatic nerve CAPs were not affected by salicylic acid-derived (aspirin; 1 mM), propionic acid-derived (ketoprofen, naproxen, ibuprofen, loxoprofen, and flurbiprofen; each 1 mM), and enolic acid-derived NSAIDs [meloxicam (0.5 mM) and piroxicam (1 mM)] [26].
The NSAIDs-induced CAP inhibition would be caused by a suppression of TTX-sensitive voltage-dependent sodium channels involved in the production of frog CAPs. Consistent with this idea, diclofenac reduced the TTX-sensitive sodium channel current amplitudes in rat dorsal root ganglion (DRG) [56] and mouse trigeminal ganglion neurons [57]. A similar sodium channel suppression caused by diclofenac has been demonstrated in myoblasts [58] and ventricular cardiomyocytes in rats [59]. Similar to diclofenac, flufenamic acid reduced the peak amplitudes of sodium channel currents in hippocampal CA1 neurons in rats [60,61,62]. Although the IC50 value (0.22 mM) needed for flufenamic acid to suppress frog sciatic nerve CAPs was close to that (0.189 mM) of sodium channel suppression in hippocampal CA1 neurons in rats [62], the IC50 value (0.94 mM) for diclofenac to inhibit CAPs was much greater than those (0.014 and 0.00851 mM in DRG neurons and myoblasts, respectively, in rats) for inhibiting sodium channels [56,58]. The rank for CAP inhibition by NSAIDs at 0.5 mM in descending order was flufenamic acid > diclofenac > indomethacin >> aspirin = naproxen = ibuprofen [26]. This order partly resembled the order for sodium channel suppression in DRG neurons (diclofenac > flufenamic acid > indomethacin > aspirin; [56]) and in cardiomyocytes (diclofenac > naproxen ≥ ibuprofen; [59]) in rats. Diclofenac at 0.3 mM reduced the TTX-resistant sodium channel current amplitudes by about 20% in trigeminal ganglion neurons in rats [63]. Flufenamic acid and tolfenamic acid at 0.1 mM reduced TTX-resistant Nav1.8 channel current amplitudes with the extents of approximately 30 and 30%, respectively [64]. Nav1.7 channel currents sensitive to TTX showed higher sensitivity to flufenamic acid and tolfenamic acid (at 0.1 mM; with the extents of approximately 60 and 70%, respectively) than TTX-resistant Nav1.8 ones [64]. Alternatively, NSAIDs inhibited the extent of increased activity produced by chemical irritation in cat corneal sensory nerve fibers; the magnitude of this inhibition was different among different types of NSAIDs [57,65]. The magnitude of NSAIDs-induced sodium channel inhibition seemed to be distinct among preparations. Concentrations needed for NSAIDs to significantly inhibit frog sciatic nerve CAPs were generally greater than those necessary for sodium channel inhibition. This result may be due to a variety of factors such as the involvement of both sodium channels and potassium channels in CAP peak amplitudes. As far as I know, it has not been reported how voltage-dependent sodium channels are affected by aceclofenac, indomethacin, etodolac, acemetacin, meclofenamic acid, and mefenamic acid. Table 1 gives a summary of the actions of NSAIDs on frog sciatic nerve fast-conducting CAPs and their IC50 values (refer to [66] as well).
More effective NSAIDs for inhibiting frog sciatic nerve CAPs have two benzene rings connected by -NH-, and hydrophilic substituents are attached to the benzene rings (refer to Figure 1(Aa,Ba) and 3(Aa,Ba,Da) in [26] for the chemical structures of diclofenac, aceclofenac, tolfenamic acid, meclofenamic acid, and flufenamic acid). Such a chemical structure is seen in local anesthetics (refer to Section 3.1) but not in the other NSAIDs [26]. Mefenamic acid, which has a hydrophobic group attached to one of the two existing benzene rings (refer to Figure 3Ca in [26]), seemed to be efficient at suppressing CAPs, although this drug was not investigated at higher concentrations owing to its low solubility (refer to above). CAPs were efficiently suppressed by 2,6-dichlorodiphenylamine and N-phenylanthranilic acid, which resemble NSAIDs in that they have two benzene rings (refer to Figs. 4Aa and 4Ba in [26]); however, they are not NSAIDs. CAPs were also suppressed by the endocrine disruptor, bisphenol A, which has two benzene rings linked to a hydrophilic group such as the hydroxyl group [67].
There is much evidence in favor of the other actions of NSAIDs depending on their chemical structures. For example, the involvement of NO-cyclic GMP-potassium channels in NSAIDs-induced analgesia depends on their chemical structures [43,68]. Nonselective cation channels expressed in the pancreas in rats were suppressed by flufenamic acid and mefenamic acid, but not by indomethacin, aspirin, or ibuprofen [69]. Diclofenac and aceclofenac inhibited TRP melastatin-3 channels in a different manner [49]. Although TRP ankyrin-1 (TRPA1) channels were depressed or activated by NSAIDs, such an activity also differed in extent among NSAIDs [70]. Furthermore, a difference was observed among NSAIDs in the activities of electron transport system or mitochondrial oxidative phosphorylation that may produce adverse side effects from NSAIDs [71].
As stated above, the concentrations for NSAIDs to inhibit frog sciatic nerve CAPs are in general much greater than those necessary for the inhibition of voltage-dependent sodium channels. Such high concentrations would be possible if NSAIDs were used in close proximity to nerve fibers. At least part of the analgesic effect of NSAIDs used as a dermatological medicine may be explained by the inhibition of nerve AP conduction by suppressed voltage-dependent sodium channels [72].
Although NSAIDs are known to exhibit a tolerance effect (reduced response to a drug after its repeated use), this effect could not be explained by the NSAID-induced peripheral nerve AP conduction inhibition revealed here, because decreased frog sciatic nerve CAP through diclofenac or aceclofenac treatment recovers to almost control levels after treatment.

2.2. Opioids

Opioids suppress glutamatergic excitatory transmission through the activation of opioid receptors expressed in the CNS such as primary-afferent fiber central terminals, which leads to antinociception ([73,74,75]; refer to reviews [76,77]). Both central and peripheral terminals of primary-afferent neurons express opioid receptors; peripheral terminal opioid receptors are known to be involved in analgesia ([78,79,80,81,82]; refer to review [83]). Opioids also have a local anesthetic effect in the PNS. Although the perineural administration of an opioid, morphine, is reported to have no effect on CAPs in the superficial radial nerve in decerebrated cats [84], APs traveling along peripheral nerve fibers are in general inhibited by opioids. For instance, opioids including fentanyl and sufentanil reduced the peak amplitudes of CAPs evoked in peripheral nerve fibers [85] and depressed the conduction of APs in peripheral nerve fibers [86]. A CAP inhibition produced by morphine in peripheral nerve fibers in mammals was sensitive to a nonspecific opioid receptor antagonist, naloxone, indicating opioid receptor activation [87]. In support of this experimental result, binding and immunohistochemical experiments have demonstrated the localization of opioid receptors in peripheral nerve fibers in mammals [88,89,90].

2.2.1. Tramadol

Tramadol [(1RS; 2RS)-2-[(dimethylamino) methyl]-1-(3-methoxyphenyl)-cyclohexanol hydrochloride] is an orally administrated opioid in the clinical setting that acts on the CNS [91]. In animals and humans, tramadol is converted to a variety of compounds including mono-O-desmethyl-tramadol (M1) through N- and O-demethylation [92]; M1 is a therapeutically active drug used to alleviate pain [91]. Among cellular mechanisms for tramadol’s antinociceptive effect, there is a μ-opioid receptor activation [93,94]. This idea is supported by the highest affinity of M1 among the metabolites of tramadol for cloned μ-opioid receptors. M1 inhibited excitatory transmission mediated by glutamate in spinal lamina II (substantia gelatinosa; SG) neurons, which play a pivotal part in the regulation of nociceptive transmission to the spinal dorsal horn from the periphery, leading to a reduced excitability of the neurons [95,96,97]. Besides such effects in the CNS, tramadol exhibits a local anesthetic effect after its intradermal application in patients ([98,99,100]; refer to review [101]). This result was consistent with in vivo studies which showed an inhibition of a spinal somatosensory evoked potential, which was produced by directly administrating tramadol into a rat sciatic nerve [102]. Tramadol added as an additive to local anesthetics has been reported to lengthen sensory block and analgesia duration [103].
CAPs recorded from the frog sciatic nerve were concentration-dependently reduced by tramadol with respect to their peak amplitude in the range of 0.2–5 mM in an irreversible manner [27]. A similar tramadol-induced CAP suppression has been demonstrated by other researchers in the frog [104] and rat sciatic nerves [105,106]. According to our analysis based on the Hill equation, the IC50 value needed for tramadol-induced frog sciatic nerve CAP amplitude reduction was 2.3 mM, which was less by approximately three-fold than that (6.6 mM) reported by Mert et al. [104] for the frog sciatic nerve. Tramadol also suppressed rat sciatic nerve CAPs (37% peak amplitude reduction at 4 mM; [105]) with an extent less than that reported by Katsuki et al. [27] for frog sciatic nerve CAPs. Tramadol action in the frog sciatic nerve was unaffected by pretreating the nerve with naloxone (0.01 mM), and (D-Ala2, N-Me-Phe4, Gly5-ol)enkephalin (DAMGO), which is an agonist for μ-opioid receptors, at 1 μM did not have any effect on frog sciatic nerve CAPs [27]. In contrast to tramadol, M1 was able to inhibit CAPs by much smaller extents (refer to below), although M1 has shown a higher affinity for μ-opioid receptors than tramadol whose chemical structure is similar to that of M1 [107]. These results indicate that opioid receptors are not involved in CAP inhibition produced by tramadol [27]. This idea is consistent with the observation that a spinal somatosensory evoked potential inhibition produced by tramadol in rat sciatic nerves in vivo was unaffected by naloxone [102]. Jaffe and Rowe [86] have also reported a naloxone-resistant nerve AP conduction inhibition produced by opioids.
Because the decreased frog sciatic nerve CAPs via tramadol treatment do not recover to control levels after treatment, repeated tramadol application may not have any effect, resulting in a tolerance. On the other hand, it is unlikely that the nerve AP conduction inhibition revealed here is related to tramadol’s addiction, as addiction is generally thought to be due to an activation of reward centers in the brain.
Although tramadol suppresses the reuptake of noradrenaline (NA) and serotonin (5-hydroxytryptamine; 5-HT) at concentrations that are sufficient for μ-opioid receptor activation [108,109], a combination of NA and 5-HT reuptake inhibitors (desipramine and fluoxetine, respectively; each at 10 μM; refer to Section 3.3) did not affect frog sciatic nerve CAPs, indicating that CAP suppression was not mediated by NA or 5-HT reuptake inhibition [27].
It is possible that the tramadol-induced CAP inhibition is mediated by the suppression of voltage-dependent sodium and potassium channels involved in AP generation. Tramadol concentration-dependently suppressed the current amplitudes of sodium channels sensitive to TTX with an IC50 value of 0.194 mM in DRG neuroblastoma hybridoma cell line ND7/23 cells [110]. Such a reduction also occurred in rat HEK293 cells expressing TTX-sensitive Nav1.2 channels; this activity had an IC50 value of 0.103 mM [111]. These values were less than the IC50 value (2.3 mM) for the inhibition of CAPs in the frog sciatic nerve [27]. Tramadol also decreased the current amplitude of delayed rectifier potassium channels (Kv3.1a type) present in NG 108-15 cells. This activity had an IC50 value of 0.025 mM, which was much smaller than the previously mentioned 2.3 mM [112]. Such IC50 values needed for tramadol to suppress CAPs, voltage-dependent sodium and potassium channels were greater than its reasonable serum concentration of approximately 2 μM in clinical practice [97,113].
On the other hand, CAPs in the frog sciatic nerve were unaffected by M1 (1–2 mM) in contrast to tramadol. Moreover, in the frog sciatic nerve exhibiting CAP inhibition by tramadol (1 mM; [27]), M1 (5 mM) suppressed CAP peak amplitudes with an extent of only 9%. Consistent with such smaller effects of M1, M1 (1 mM) did not block APs conducting on rat primary-afferent fibers when its effect on excitatory postsynaptic currents evoked by stimulating the dorsal root was investigated by use of whole–cell patch clamp recordings from SG neurons in spinal cord slices [96]. Interestingly, -OCH3 is bound to the benzene ring for tramadol, whereas it is -OH for M1; thus, the methyl group is located on tramadol, but not on M1 (refer to Figure 5a in [27]). In conclusion, the distinction in CAP inhibition between tramadol and M1 could be explained by the difference in chemical structure.

2.2.2. Other Opioid-Related Compounds (Morphine, Codeine, and Ethylmorphine)

In order to know whether the structure–activity relationship between tramadol and M1 is applied to other opioids, it was examined how frog sciatic nerve CAPs are affected by morphine, codeine (where -OCH3 is present instead of -OH in morphine), and ethylmorphine (where the -OH of morphine is replaced by -OCH2CH3; refer to Figure 7A in [28] for their chemical structures). Morphine at 5 mM concentration reduced the CAP peak amplitude with an extent of 15% and morphine activity was concentration-dependent; codeine (5 mM) suppressed the CAP peak amplitude with an extent of 30%. Ethylmorphine inhibited CAPs in a manner more effective than morphine and codeine (inhibition at 5 mM: 61%). Ethylmorphine activity was concentration-dependent and had an IC50 value of 4.6 mM. The activities of morphine, codeine, and ethylmorphine were resistant to naloxone (0.01 mM). Naloxone (1 mM) itself reduced the CAP peak amplitudes by 9% while having no effect on morphine activity [28]. These results indicate that the CAP inhibitions produced by opioids were not mediated by opioid receptors, an observation similar to those of mammalian peripheral nerves [85,86,114]. On the contrary, Hunter and Frank [115] have reported a naloxone-sensitive CAP inhibition in the frog sciatic nerve. The order of the CAP peak amplitude reduction produced by opioids is ethylmorphine > codeine > morphine, indicating that the extent of CAP inhibition increases with increasing number of -CH2. The observed results are consistent with the relationship between tramadol and M1 mentioned above. It is of interest to note that morphine, codeine, and ethylmorphine are absolutely different in chemical structure from tramadol and M1 [116]. Because increasing the number of -CH2 groups present in an opioid increases its lipophilicity, an interaction between lipophilic opioids and ion channels has been suggested to play a crucial part in the inhibition of nerve AP conduction, as demonstrated for local anesthetics [117,118]. Structure–activity relationships similar to that of opioids have been demonstrated for other chemicals. The potency of CAP inhibition in the rat sciatic nerve decreased in the order of isopropylcocaine > cocaethylene > cocaine [119]. Interestingly, the affinity of opioids for μ-opioid receptors decreased in the order of morphine, codeine, and ethylmorphine [120]; this order is reversed to that of CAP suppression. This result supports the idea that CAP inhibition produced by opioids in the frog sciatic nerve is not thanks to opioid receptor activation.
Although opioids including morphine, codeine, and ethylmorphine are known to have a tolerance effect, this could not be explained by their inhibitory effects on nerve AP conduction, because their opioids reversibly suppress frog sciatic nerve CAPs [28].
The suppression of CAP, similar to that seen in the frog sciatic nerve, has also been shown in the mammalian peripheral nerve, albeit to varying extents. A frog sciatic nerve CAP peak amplitude reduction (by about 30%) produced by codeine (5 mM) was much smaller than the reduction observed in the rat phrenic nerve (by approximately 70%), whereas there was not so large a distinction in morphine (5 mM) action (reduction by approximately 10%) between the two preparations. Morphine sensitivity was lower in the frog sciatic nerve compared with rabbit and guinea pig vagus nerves, whose CAP peak amplitudes were decreased with the extenst of 20–32% at 0.5 mM [87]. APs obtained by intracellular recordings from rat DRG neurons with Aα/β myelinated primary-afferent fibers were also inhibited by opioids in a decreasing order of ethylmorphine, codeine, and morphine (whose IC50 values were 0.70, 2.5, and 2.9 mM, respectively) with respect to AP peak amplitude decrease. These AP inhibitions were also resistant to naloxone (0.01 mM) [121].
AP conduction in peripheral nerve fibers is inhibited by a variety of drugs such as narcotics, antiepileptics, local anesthetics, alcohols, and barbiturates, suggesting that the drugs may interact with membrane bilayers in a nonspecific manner [122]. However, the chemical structure-dependent suppression of CAPs by opioids as described above indicates that opioids act on membrane proteins including voltage-dependent sodium and potassium channels [123]. In support of this idea, morphine inhibited the peak sodium channel currents and steady-state potassium channel currents recorded from frog sciatic nerve single myelinated fibers, resulting in the prolongation of APs [124]. The intracellular application of morphine resulted in the reduction of voltage-dependent sodium and potassium channel current amplitudes in squid giant axons [125]. Bath application of morphine reduced the peak amplitudes of sodium channel currents, sensitive to TTX, recorded from ND7/23 cells (neuroblastoma x DRG neuron hybrid cell line); this activity had an IC50 value of 0.378 mM [110], whereas morphine at 1 mM did not affect TTX-sensitive Nav1.2 channels located in HEK293 cells [111]. Supporting the idea regarding ion channel inhibition, the opioid meperidine was prescribed for a blockade of AP conduction and in turn antinociception; meperidine depressed sodium channels in a fashion resembling that of lidocaine [126]. Table 1 gives a summary of the inhibitory actions of opioids on frog sciatic nerve fast-conducting CAPs together with their IC50 values (refer to [66] as well).
In clinical practice, much of the pain relief from opioids is due to the administration of centrally permeable opioids into the systemic circulation, which results in opioid actions in both the PNS and CNS, leading to analgesia [127]. Opioids applied into the nerve sheath can also alleviate pain (for instance, refer to [128]). Opioids that are centrally administered can act on the PNS as well as the CNS, because opioids are moved from the brain to the periphery by an action of P-glycoproteins [129]. The subcutaneous administration of blood–brain barrier-impermeable N-methyl-morphine produced antinociception in an acetic acid-writhing mice model [78]. Brain-impermeable opioid loperamide subcutaneously administered showed an antinociceptive effect in a formalin test on rats [80]. Thus, nerve AP conduction inhibition produced by opioids might contribute to local antinociception after the perineural application of opioids in the periphery (for example, refer to [130]), which may lead to a direct action of opioids at high doses on peripheral nerves. Peripherally-applied codeine might have a nerve AP conduction inhibitory effect similar to morphine, because codeine is converted to morphine by O-demethylation in humans and animals ([131,132]; refer to review [127]).
Although opioids exhibit side effects such as tolerance and addiction, these effects appear to be mainly due to the action on synaptic transmission in the CNS rather than nerve AP conduction. This is because synaptic transmission can be plastically changed in efficacy, but this is not the case for nerve AP conduction.

3. Actions of Analgesic Adjuvants on Nerve AP Conduction

3.1. Local Anesthetics

Local anesthetics inhibit both voltage-dependent sodium and potassium channels ([117]; refer to reviews [123,133,134]). Owing to this inhibition, local anesthetics have been used to alleviate neuropathic pain in the hope of suppressing nerve AP conduction in animals [135,136] and humans [137,138,139,140], although it is possible that other effects such as the modulation of neurotransmitter receptors, toll-like receptors, and TRP channels are also involved in analgesia [133]. Various types of local anesthetics have been reported to open TRPA1 channels located in primary-afferent neuron central terminals in the SG of the rat spinal dorsal horn [141,142] and TRPA1 and TRP vanilloid-1 (TRPV1) channels in rodent DRG neurons [143,144].

3.1.1. Amide-Type Local Anesthetics

Frog sciatic nerve CAPs were reversibly and concentration-dependently reduced in terms of peak amplitude by the amide-type local anesthetic lidocaine, which blocks nerve AP conduction [104,105,106,134]. When examined in the range of 0.1–2 mM, lidocaine’s IC50 value was 0.74 mM [28]. This value was greater than that observed for the voltage-dependent sodium channel current amplitude reduction (0.204 mM) while being smaller than the voltage-dependent potassium channel current amplitude reduction (1.118 mM) in sciatic nerve fibers in the toad Xenopus laevis [118]. The peak amplitudes of sodium channels resistant to TTX in rats were concentration-dependently reduced by lidocaine; an IC50 value for this activity was 0.073 mM [145], which was 10 times less than that for the reduction of frog sciatic nerve CAP peak amplitudes. At least part of antinociception produced by systemically applied lidocaine in humans [146] may be attributed to its inhibitory effect on nerve AP conduction.
A resembling reversible CAP peak amplitude reduction was caused by another amide-type local anesthetic, ropivacaine, which exhibits a longer duration of action than lidocaine in blocking nerve conduction of APs ([147]; refer to review [148]). Ropivacaine activity was concentration-dependent over concentrations ranging from 0.01 to 1 mM; this activity had an IC50 value of 0.34 mM [27]. The frog sciatic nerve ropivacaine-induced CAP amplitude reduction was almost comparable to that observed in rabbit vagus nerve A-fibers (approximately 30% at 0.2 mM) [149]. Frog sciatic nerve IC50 values for lidocaine and ropivacaine, which are 0.74 and 0.34 mM, respectively, were not significantly distinct from those for fast-conducting CAP peak amplitude reduction in the rat sciatic nerve (0.28 mM for both lidocaine and ropivacaine) [150]. Moreover, an amide-type local anesthetic, prilocaine, was also shown to reversibly and concentration-dependently reduce frog sciatic nerve CAP peak amplitudes. When examined at concentrations ranging from 0.01 to 5 mM, prilocaine’s IC50 value was 1.8 mM [67].
There are levobupivacaine and its racemic, bupivacaine, which are amide-type local anesthetics; the former has a lower cardiovascular risk and CNS toxicity than the latter ([151]; refer to review [152]). CAP peak amplitudes in the frog sciatic nerve were reversibly and concentration-dependently reduced by levobupivacaine. When examined at concentrations ranging from 0.05 to 1 mM, levobupivacaine’s IC50 value was 0.23 mM [30]. This IC50 value was close to a previously reported value (0.22 mM) for a tonic levobupivacaine-induced inhibition of frog sciatic nerve CAPs [151], and it was also close to a previously reported value (0.264 mM) for a tonic levobupivacaine-induced suppression of voltage-dependent sodium channel currents, which were recorded at the holding potential of −100 mV in GH-3 neuroendocrine cells [153]. As shown in the previous study [151], the levobupivacaine-induced CAP amplitude reduction in the frog sciatic nerve was less than that of bupivacaine (their extents at 0.5 mM were 45 and 76%, respectively [30]). This frog sciatic nerve bupivacaine activity was less than the observed values for toad Xenopus laevis sciatic nerve fiber sodium channels (IC50 = 0.027 mM) [118], sodium channels sensitive to TTX in ND7/23 cells (IC50 = 0.178 mM) [154], and rat clonal pituitary GH3 cell sodium channels (IC50 = 0.190 mM) [155]. Voltage-dependent potassium channels in sciatic nerve fibers in the toad Xenopus laevis were also suppressed by bupivacaine; this sensitivity (IC50 = 0.092 mM) was less than that for sodium channels [118].

3.1.2. Ester-Type Local Anesthetics

As a classic ester-type local anesthetic, there is a compound (cocaine) isolated from the coca plant Erythroxylon coca, which has been long known to suppress nerve AP conduction ([114,156,157]; refer to review [158]). Frog sciatic nerve CAP peak amplitudes were reversibly and concentration-dependently reduced by cocaine. When examined at concentrations ranging from 0.01 to 2 mM, an IC50 value for cocaine activity was 0.80 mM [28]. This value was similar to that obtained for lidocaine in the frog sciatic nerve (0.74 mM) [27]. The cocaine’s IC50 value was about four times larger than that observed for the rat phrenic nerve (approximately 0.2 mM) [114]. Although cocaine (40 μM) reduced mouse phrenic nerve CAP peak amplitudes by 26% [157], such a reduction in the frog sciatic nerve occurred at a concentration of approximately 300 μM [28]. There was a similar extent of CAP peak amplitude reduction by cocaine in the frog and rat sciatic nerve (30% at 0.5 mM in frogs; 40% at 0.375 mM in rats; refer to [119]). There is ample evidence for suppression of voltage-dependent sodium channels by cocaine (for example, refer to [119,159,160]); a cocaine (0.05 mM)-induced tonic (TTX-resistant) Nav1.5 channel current amplitude reduction was about 70% [160]. Cocaine and lidocaine suppressed voltage-dependent sodium channels in a competitive manner [161]. Cocaine also inhibited delayed rectifier potassium channels in central snail neurons [162].
Another ester-type local anesthetic, procaine, [163] also reduced the peak amplitudes of CAPs in the frog sciatic nerve in a reversible and concentration-dependent manner. When examined at concentrations ranging from 0.1 to 5 mM, procaine’s IC50 value was 2.2 mM [164]. This value was close to those (2–5 mM) that other researchers [165,166] reported in the same preparation and to the value obtained in the rat sciatic nerve (about 1 mM) [165]. Moreover, a ratio (2.2 mM/0.74 mM) of procaine’s IC50 value to that of lidocaine [27] in reducing frog sciatic nerve CAP amplitudes was comparable to the ratio (0.53%/0.14%) of procaine concentration necessary to block motor nerve AP conduction by 50%, to lidocaine’s one in rats [167]. In addition, procaine activity in the frog sciatic nerve was 37 times less than that for voltage-dependent sodium channel current amplitude reduction in sciatic nerve fibers in the toad Xenopus laevis (IC50 = 0.060 mM) [118]. Procaine also suppressed voltage-dependent potassium channels in this preparation with an IC50 value (6.303 mM) which was larger than that obtained for sodium channels [118].
Benzocaine (ethyl 4-aminobenzoate), which is an ester-type local anesthetic, is used for not only topical anesthesia in clinical practice [168] but also amphibian anesthesia ([169]; refer to reviews [170,171]). Frog sciatic nerve CAP peak amplitudes were reversibly and concentration-dependently reduced by benzocaine. When examined at concentrations ranging from 0.01 to 2 mM, an IC50 value for benzocaine activity was 0.80 mM (73% inhibition at 1 mM [29]). The rat sciatic nerve exhibited a similar benzocaine-induced CAP inhibition (37% inhibition at 1.3 mM [105]). The benzocaine activity was similar to those observed with cocaine and lidocaine.
Another ester-type local anesthetic, tetracaine, was also shown to reversibly and concentration-dependently reduce frog CAP peak amplitudes. When examined at concentrations ranging from 0.0005 to 0.05 mM, tetracaine’s IC50 value was 0.014 mM [32]. This value is similar to that of frog sciatic nerve fibers (0.0063 mM) as reported by Starke at al. [172] and also to that obtained for rabbit A nerve fibers (0.009 mM) [173]. On the other hand, the frog sciatic nerve tetracaine activity was 19 times smaller than that present in suppressing voltage-dependent sodium channel currents in toad Xenopus laevis sciatic nerve fibers (IC50 = 0.0007 mM) [118]. In this preparation, voltage-dependent potassium channel current amplitudes were also reduced by tetracaine; an IC50 value (0.946 mM) for this activity was much greater than for the sodium channels [118]. Tetracaine suppressed both frog sciatic nerve CAPs and toad Xenopus laevis sodium channels much more effectively than procaine, lidocaine, and bupivacaine.
The peak amplitudes of frog sciatic nerve CAPs were also attenuated by pramoxine that is a non-amide- and non-ester-type local anesthetic. This inhibitory action of pramoxine was concentration-dependent over concentrations ranging from 0.001 to 1 mM; this activity had an IC50 value of 0.21 mM and subsided with a slow time course after pramoxine washout [67]. Table 1 summarizes the inhibitory actions of local anesthetics on frog sciatic nerve fast-conducting CAPs together with their IC50 values (see [66] as well).
In clinical practice, local anesthetics locally administered to peripheral nerves may be repeatedly effective without being tolerated, because both amide-type and ester-type local anesthetics reversibly inhibit frog sciatic nerve CAPs.

3.2. Antiepileptics

Antiepileptics exhibit a variety of actions, such as glutamate receptor inhibition, GABAA receptor activation, and voltage-dependent sodium and calcium channel inhibition [13,174]. Antiepileptics are well-known to inhibit neuropathic pain (for example, refer to [175]). As indicated by the action on sodium channels, it is possible that the neuropathic pain alleviation is due to nerve AP conduction inhibition.
The CAP peak amplitudes of frog sciatic nerves were attenuated by lamotrigine (3,5-diamino-6-(2,3-dichlorophenyl)-1,2,4-triazine) that is a phenyltriazine derivative. Lamotriginine is reported to depress the voltage-dependent sodium channels [176] and also to attenuate a neuropathic pain accompanied by a cerebrovascular accident or diabetic polyneuropathy [7]. The sciatic nerve lamotrigine activity was in part reversible and concentration-dependent over concentrations ranging from 0.02 to 0.5 mM, and an IC50 value for this activity was 0.44 mM [30]. This value was similar to the IC50 value (0.641 mM at −90 mV) for lamotrigine to inhibit human brain type IIA sodium channels, sensitive to TTX, present in Chinese hamster ovary cells [176]. A similar CAP amplitude reduction was reported for carbamazepine (5H-dibenz[b,f]azepine-5-carboxamide; [30]), which is an iminostilbene derivative, is distinct in chemical structure from lamotrigine, and inhibits voltage-dependent sodium channels [177]. Carbamazepine has been reported to effectively attenuate trigeminal neuralgia [178,179]. Unlike the case of lamotrigine, the carbamazepine-induced CAP inhibition in the frog sciatic nerve was completely reversible. This carbamazepine activity was concentration-dependent over concentrations ranging from 0.05 to 1 mM, and an IC50 value for this activity was 0.50 mM [30]. Carbamazepine and lamotrigine attenuated sodium channel currents in N4TG1 mouse neuroblastoma cells with similar IC50 values [180], an observation being consistent with the experimental result that the two antiepileptics had comparable efficacies in frog sciatic nerve CAP amplitude reduction.
Oxcarbazepine (10,11-dihydro-10-oxo-5H-dibenz[b,f]azepine-5-carboxamide; [181]), which differs from carbamazepine in that it has a keto substitution at the 10,11 position of the dibenzazepine nucleus, reduced frog sciatic nerve CAP peak amplitudes; this efficacy was less than carbamazepine’s one [30]. Oxcarbazepine is known to effectively relieve pains accompanied by diabetes [7] and trigeminal nerve injury [179]. Oxcarbazepine activity was partially reversible and concentration-dependent over concentrations ranging from 0.02 to 0.7 mM. The extent of CAP amplitude reduction (40%) by oxcarbazepine (0.7 mM) was slightly less than that (57%) of carbamazepine (0.7 mM). Each of lamotrigine, carbamazepine, and oxcarbazepine (all 0.5 mM) augmented the threshold required for inducing frog sciatic nerve CAPs [30]. Consistent with this experimental result, their antiepileptics reduced voltage-dependent sodium channel current amplitudes by shifting their steady-state inactivation to a more negative membrane potential [180,182,183]. Frog sciatic nerve CAP amplitude reduction produced by oxcarbazepine at 0.5 mM (20%; [30]) was much less in extent than that for the inhibition of sodium channel currents sensitive to TTX in differentiated NG108-15 neuronal cells (IC50 = 3.1 μM; [182]). Consistent with the observation that oxcarbazepine exhibited a smaller frog sciatic nerve CAP inhibition than carbamazepine, oxcarbazepine was less effective than carbamazepine in attenuating maximal electroshock-induced seizures in rats [181].
Another antiepileptic, phenytoin (hydantoin derivative, 5,5-diphenylhydantoin, which suppresses voltage-dependent sodium channels [184] and attenuates paroximal attack accompanied by trigeminal neuralgia [179]), concentration-dependently inhibited frog sciatic nerve CAPs by a small extent over concentrations ranging from 0.01 to 0.1 mM; the magnitude was only 15% at 0.1 mM [30]. The frog sciatic nerve phenytoin activity was smaller than those of rat cortical and human type IIA sodium channels, the amplitudes of which were decreased by 60–90% by phenytoin (0.1 mM) at −60 mV [176,184]. Unlike frog sciatic nerve CAPs, voltage-dependent sodium channels were suppressed by phenytoin with an IC50 value similar to lamotrigine in N4TG1 mouse neuroblastoma cells [182]; phenytoin, lamotrigine, and carbamazepine reportedly bound to a common site present on sodium channels in hippocampal CA1 neurons in rats [183]. The magnitude of a sensitivity of voltage-dependent sodium channels to phenytoin appeared to differ among different types of the channel. In support of this idea, the magnitudes of phenytoin activities were distinct among human Nav1.1, Nav1.2, Nav1.3, and Nav1.4 α-subunits, sensitive to TTX, located in HEK293 cells [185]. Moreover, the properties and accessibilities of sodium channels differed between myelinated nerve fibers in frogs and rats [186].
Antiepileptics that can inhibit CAPs resemble NSAIDs in chemical structure, judging by the fact that two unsaturated six-membered rings are contained in lamotrigine, carbamazepine, and oxcarbazepine (for the chemical structures of the three antiepileptics, refer to Figures 1a, 2aA, and 2bA in [30]). Because the inhibitory effects of carbamazepine and diclofenac on voltage-dependent sodium channels occlude each other, they seem to have a common or closely related binding site [61].
In addition, frog sciatic nerve CAPs were unaffected by the other antiepileptics at concentrations as high as 10 mM [30]. Among them, there are gabapentin (1-(aminomethyl)cyclohexaneacetic acid, which has a chemical structure similar to GABA and attenuates pain persisting after herpes zoster [179]), topiramate (2,3:4,5-bis-O-(1-methylethylidene)-β-D-fructopyranose sulfamate, which relieves various neuropathic pains such as neuralgia involving the intercostal and trigeminal nerves [13]), and sodium valproate (2-propylpentanoic acid sodium salt, which alleviates neuropathic pain caused by diabetes [13]). The weak inhibitory effects of gabapentin and sodium valproate on frog sciatic nerve CAPs were seen with human type IIA sodium channels [176]. The human sodium channels were hardly affected by gabapentin at concentrations of less than 3 mM [176]. The antinociceptive action of gabapentin would be attributed to the fact that it binds to voltage-dependent calcium channel α2δ-1 subunit, resulting in an inhibited influx of calcium ions into nerve terminals and in turn an attenuation of the release of neurotransmitters from them [187]. Gabapentin may also interrupt an interaction between N-methyl-D-aspartate (NMDA) receptor (which is one subtype of glutamate receptors) channels and voltage-dependent calcium channel α2δ-1 subunit in postsynaptic neurons [188]. As distinct from the frog sciatic nerve, TTX-sensitive sodium channels in rat cerebellar granule cells were suppressed by topiramate; this activity had an IC50 value of 0.0489 mM [189]. Such a difference between the sciatic nerve and cerebellar granule cells would be possibly due to a distinction in topiramate sensitivity among different sodium channel types or phosphorylated states [190]. Antinociception produced by sodium valproate and topiramate has been attributed to other mechanisms including an increase in GABAA receptor response [191,192]. An inhibition of glutamate receptors would also possibly contribute to topiramate- and lamotrigine-induced antinociceptions. This is because topiramate depresses GluK1 (GluR5) kainate receptors (one subtype of glutamate receptors) in basolateral amygdala neurons [193] and lamotrigine suppresses α-amino-3-hydroxy-5-methyl-4-isoxazole propionate (AMPA) receptors (another subtype of glutamate receptors) in dentate gyrus granule cells in rats [194]. Table 1 gives a summary of the actions of antiepileptics on frog sciatic nerve fast-conducting CAPs and their IC50 values (see [66] as well).
Antiepileptics that can inhibit frog sciatic nerve CAPs seemed to exhibit antinociceptive actions in a persistent pain model. Intraperitoneally administrated lamotrigine, carbamazepine, and oxcarbazepine caused analgesia in the second phase of the formalin test (which reflects inflammation occurring 15–20 min after formalin injection) but phenytoin, topiramate, and sodium valproate had no effect in rats [195,196]. The antinociception produced by antiepileptics appeared to be partly due to their nerve AP conduction inhibitory actions. The plasma concentrations of lamotrigine and carbamazepine clinically prescribed to relieve epilepsy were less than 12 μM and 20–50 μM, respectively [197,198]. These values were much smaller than the IC50 values needed for them to inhibit frog sciatic nerve CAPs.
In clinical practice, antiepileptics locally administered to peripheral nerves may be repeatedly effective without being tolerated, as the inhibitory actions of antiepileptics on frog sciatic nerve CAPs are partial but reversible.

3.3. Antidepressants

Antidepressants are thought to alleviate pain by activating the descending analgesic pathway composed of 5-HT- or NA-containing nerve fibers to the spinal dorsal horn from brainstem by a suppression of their neurotransmitters’ reuptake [199,200], involvement of α adrenoceptors, H1-histamine, 5-HT, opioid and muscarinic acetylcholine receptors ([11,201,202,203,204]; refer to reviews [205,206]), and the suppression of voltage-dependent calcium [207,208], NMDA receptor [209,210,211,212], and purinergic P2X4 receptor (one subtype of ionotropic P2X receptors activated by ATP; [213]) channels, all of which are involved in modulating synaptic transmission. Moreover, an inhibition of neuroimmune mechanisms accompanying nerve injury may be involved in the pain alleviation produced by antidepressants [214].
Duloxetine, which is 5-HT and NA reuptake inhibitor (SNRI), ([215,216,217]; refer to review [218]) partially but reversibly inhibited frog sciatic nerve CAPs. Duloxetine activity was concentration-dependent over concentrations ranging from 0.001 to 2 mM, and an IC50 value for this activity was 0.23 mM [31]. A resembling CAP inhibition was caused by fluoxetine that is a selective inhibitor of 5-HT reuptake (SSRI) ([201,202]; refer to reviews [206,219]). Fluoxetine-induced CAP peak amplitude reduction was partially but reversible, concentration-dependent over concentrations ranging from 0.05 to 5 mM, and an IC50 value for this activity was 1.5 mM. Thus, fluoxetine was less effective in inhibiting CAPs than duloxetine [31].
Typical tricyclic antidepressants, amitriptyline and desipramine, which have tertiary and secondary amine structures, respectively [200,203,220,221], also inhibited frog sciatic nerve CAPs. Amitriptyline reduced CAP peak amplitudes over concentrations ranging from 0.001 to 1 mM, and an IC50 value for this activity was 0.26 mM. Desipramine also showed a similar effect over concentrations ranging from 0.1 to 5 mM; this activity had an IC50 value of 1.6 mM [31]. Thus, amitriptyline inhibited CAPs 6-fold more effectively than desipramine. Consistent with this finding, amitriptyline blocked AP conduction in the rat sciatic nerve [155]. Like tricyclic antidepressants, maprotiline, which is a tetracyclic antidepressant [220], also exhibited an inhibitory action on frog sciatic nerve CAPs in a partial but reversible fashion. Maprotiline activity was concentration-dependent over concentrations ranging from 0.2 to 5 mM, and an IC50 value for this activity was 0.95 mM [31]. Trazodone, a 5-HT2 receptor antagonist and reuptake inhibitor (SARI), is a non-SNRI, non-SSRI, non-tricyclic, and non-tetracyclic antidepressant ([222,223,224,225]; refer to review [226]). Trazodone reduced frog sciatic nerve CAP peak amplitudes in a partial but reversible fashion at concentrations ranging from 0.2 to 2 mM which is the maximum soluble concentration. The magnitude of trazodone (1 mM)-induced CAP peak amplitude reduction was about 50% [31].
The antidepressants-induced CAP inhibition would be owing to an attenuation of voltage-dependent sodium channels sensitive to TTX that are involved in frog sciatic nerve CAP generation. Consistent with this idea, voltage-dependent sodium channels were inhibited by duloxetine [227,228], fluoxetine [229], amitriptyline [227,229,230,231,232,233,234,235], desipramine, and maprotiline [236]. Voltage-dependent sodium channels sensitive to TTX, which are located in adrenal chromaffin cells in bovines, were suppressed by amitriptyline (with an IC50 value of 0.0202 mM), fluoxetine (0.02 mM; 62% amplitude reduction), desipramine (0.02 mM; 50% reduction), and trazodone (0.1 mM; 20% reduction; [229]). Amitriptyline also reduced the peak amplitudes of sodium channel currents in clonal pituitary GH3 cells in rats; this activity had an IC50 value of 0.0398 mM [155]. These efficacies for sodium channel inhibition were much greater than those obtained for frog sciatic nerve CAPs. Moreover, an IC50 value (0.0221 mM) for the inhibition of Nav1.7 channel currents, sensitive to TTX, produced by duloxetine was approximately 10 times less than that (0.23 mM) for inhibiting CAPs in the frog sciatic nerve [228]. The efficacy sequence that maprotiline more effectively depresses CAPs than fluoxetine was the same as that reported for Nav1.7 channels, where IC50 values for inhibiting the channels by maprotiline, fluoxetine, desipramine, and amitriptyline were 0.028, 0.074, 0.024, and 0.085 mM, respectively [236]. The experimental result that amitriptyline and duloxetine inhibited frog sciatic nerve CAPs with a similar IC50 value was the same as that reported for cardiac-type sodium channel suppression [227]. The peak amplitudes of sodium channel (possibly Nav1.8 channel)currents, resistant to TTX, in trigeminal ganglion neurons in rats were also reduced by amitriptyrine; this activity had an IC50 value of 0.00682 mM [237]. In terms of chemical structure, typical local anesthetics have hydrophilic and hydrophobic moieties with an intermediate amide or ester linkage between them [238], while all of the antidepressants, except for trazodone, examined in the frog sciatic nerve, consist of a hydrophilic amine group and a hydrophobic moiety containing benzene rings, both of which are connected by a straight hydrocarbon chain (refer to Figure 1 in [31] for their antidepressants’ chemical structures). Such chemical structures may play a pivotal role in inhibiting sodium channels. Table 1 gives a summary of IC50 values for antidepressants to inhibit fast-conducting CAPs in the frog sciatic nerve (see [66] as well).
The antidepressants examined in the frog sciatic nerve are clinically used to alleviate chronic pain [10,11,217,218,221,239,240] and attenuate neuropathic pain in animal models. For instance, duloxetine attenuated tactile allodynia (where pain is caused by a stimulus that is normally painless) and heat hyperalgesia (where the sensitivity to painful stimuli is abnormally increased) in neuropathic pain models in rats [215]. Fluoxetine caused analgesia in diabetic neuropathic pain models produced by streptozotocin in mice [201]. Amitriptyline and desipramine effectively attenuated pain in patients suffering from diabetic neuropathy [200]. Maprotiline suppressed neuropathic pain in rats undergoing chronic constriction injury to the sciatic nerve [241]. Trazodone depressed hyperalgesia produced in chronic constriction injury models in rats [222]. The plasma concentrations of duloxetine, fluoxetine, amitriptyline, desipramine, maprotiline, and trazodone prescribed for the treatment of depression and neuropathic pain in clinical practice are 0.09–0.3, 0.3–1.6, 0.36–0.90, 0.47–1.1, 0.72–1.4, and 2.2–4.3 μM, respectively [205,228]. These concentration values were much smaller than the IC50 values needed for them to inhibit frog sciatic nerve CAPs. The antidepressants may only produce an analgesia when locally applied to the nerve.
In clinical practice, antidepressants locally administered to peripheral nerves may be repeatedly effective without being tolerated, as the inhibitory actions of antidepressants on frog sciatic nerve CAPs are partial but reversible.

3.4. Adrenoceptor Agonists

The epidural and intrathecal administration of α2 adrenoceptor agonists such as clonidine and DEX [242] results in antinociception in animals [243,244,245] and humans [246]. This is possibly due to the inhibition produced by the agonists of excitatory transmission mediated by glutamate in spinal superficial dorsal horn neurons [247]. Administration of α2 adrenoceptor agonists and local anesthetics for the purpose of spinal anesthesia prolongs peripheral nerve conduction block in animals [248,249,250,251] and humans ([252,253,254,255,256,257,258,259]; refer to review [260]). This is possibly due to a local vessel contraction produced by the agonists, leading to decreased removal of the anesthetics from the subarachnoid space [261,262]. Furthermore, α2 adrenoceptor agonists attenuate nerve AP conduction and thus have a local anesthetic effect, leading to an enhanced effect of local anesthetic [263]. For instance, clonidine produced both a suppression of excitatory transmission in rat spinal SG neurons [264,265] and a blockade of peripheral nerve AP conduction [172,263,266]; the latter effect required much higher concentrations of clonidine than the former effect. Similar to clonidine, DEX reportedly inhibited excitatory transmission in rat SG neurons [267]. When DEX or clonidine, combined with lidocaine, was intracutaneously applied into the back of guinea pigs, the local anesthetic effect of lidocaine was enhanced [268]. Local wound infiltration with DEX together with another local anesthetic, bupivacaine, more effectively attenuated pain that occurred after surgery than bupivacaine alone in humans [269]. In addition to clonidine, DEX has a possible inhibitory action on nerve AP conduction, as DEX is reported to suppress voltage-dependent sodium channels [145] (refer to below).
The CAP peak amplitudes of the frog sciatic nerve were reduced by DEX in a concentration-dependent and reversible fashion. When examined at concentrations ranging from 0.01 to 1 mM, DEX’s IC50 value was 0.40 mM [32]. The DEX activity was resistant to the α2-adrenoceptor antagonists, yohimbine and atipamezole ([247,270,271,272]; refer to reviews [273,274]), although DEX exhibited a high affinity for α2 adrenoceptors [242]. This result indicates no involvement of α2 adrenoceptors in the DEX activity [32]. CAP peak amplitude reductions were also produced by another α2-adrenoceptor agonist, oxymetazoline, which is more selective to α2A than α2B and α2C adrenoceptors [273,275]; clonidine also presented CAP peak amplitude reductions in a manner insensitive to yohimbine. Oxymetazoline concentration-dependently and reversibly reduced the CAP peak amplitude, and an IC50 value for this activity was 1.5 mM. Clonidine at 2 mM reduced the CAP peak amplitude by about 20% [32]. The extent of this clonidine activity was distinct from the results (CAP amplitude reduction of 80% at 0.3 mM) obtained by Starke et al. [172] using the same frog sciatic nerve, although the reason why this difference occurs is not known. In addition, a variety of adrenoceptor agonists, adrenaline, NA, phenylephrine (α1-adrenoceptor agonist), and isoproterenol (a β-adrenoceptor agonist) at 1 mM did not affect CAPs in the frog sciatic nerve [32]. A similar CAP peak amplitude reduction caused by clonidine has been demonstrated in the sciatic nerve in rats. Thus, CAPs derived from primary-afferent Aα and C fibers contained in the sciatic nerve in rats were found to be suppressed by clonidine; IC50 values in their fibers were 2.0 and 0.45 mM, respectively [266].
The α2-adrenoceptor agonists-induced CAP inhibition would be mediated by a suppression of voltage-dependent sodium and potassium channels that are involved in AP generation. DEX reportedly reduced the voltage-dependent sodium channel current amplitude in DRG neurons in a manner resistant to yohimbine in rats, although this type of sodium channels was insensitive to TTX [145]. The IC50 value (0.058 mM) for this DRG neuron DEX activity in rats was approximately ten times less than that (0.40 mM) required for sciatic nerve CAP suppression in frogs. The peak amplitude of sodium channel current insensitive to TTX in rats was also reduced by clonidine; this activity had an IC50 value of 0.26 mM [145]. The peak amplitudes of TTX-sensitive sodium channel currents in ND7/23 cells were reduced by clonidine; an IC50 value for this reduction was 0.824 mM [154]. It has been demonstrated in NG108-15 neuronal cells that the peak amplitudes of delayed rectifier potassium channel currents were reduced by DEX; an IC50 value for this activity was 0.0046 mM, while the peak amplitudes of sodium channel currents sensitive to TTX were attenuated by approximately 20% by DEX (0.01 mM) in a manner resistant to yohimbine [276]. These differences in drug potency may be caused by a distinction in either sodium channel types or animal species. Table 1 demonstrates the actions of adrenoceptor agonists on frog sciatic nerve fast-conducting CAPs together with their IC50 values (refer to [66] as well).
In clinical practice, DEX administration results in producing analgesia/sedation and decreasing heart rate, cardiac output, and memory; different plasma concentrations of DEX are effective in each of these effects [277]. In patients, sedation is rapidly induced by administering at a rate of 0.2–0.7 mg·kg−1·hr−1 (intravenous) [242]. In intramuscular applications in cats, 40 mg·kg−1 is the dose used usually to produce analgesia/sedation [278]. The concentrations of DEX necessary for frog sciatic nerve AP conduction inhibition are more than 1000 times higher than those of DEX used as α2 adrenoceptor agonist, as plasma levels of DEX in clinical use are less than 0.05 μM (refer to [277]). Therefore, the potency of DEX in blocking nerve AP conduction does not depend on the usage of DEX for analgesia/sedation. α2-Adrenoceptor agonists such as DEX, together with a local anesthetic, have been applied for prolonging the duration of AP conduction block in peripheral nerve fibers [249,252,253,254,255,279]. This effect is possibly due to a local vasoconstriction leading to slowing local anesthetic absorption and/or a direct nerve AP conduction suppression produced by α2 adrenoceptor agonists [256]. The latter mechanism would be the α2 adrenoceptor agonists-induced nerve CAP inhibition, as stated above. This action becomes meaningful when their topical administration on nerves is considered but is not related to analgesia/anesthesia caused by their systemic application. Certain chemical structures of their α2 adrenoceptor agonists (refer to [32]) may play a crucial role in the production of nerve AP conduction blockage. DEX locally administered to peripheral nerves may be repeatedly effective without being tolerated, as DEX reversibly inhibits frog sciatic nerve CAPs.

4. Comparison in Nerve AP Conduction Inhibition among Analgesics and Analgesic Adjuvants

As noted from Table 1, some analgesic adjuvants had similar IC50 values for frog sciatic nerve CAP inhibitions. For example, antidepressants had IC50 values resembling those of some of local anesthetics, antiepileptics, and α2-adrenoceptor agonists. The IC50 values of duloxetine and amitriptyline, which are 0.23 and 0.26 mM, respectively, resemble the values that ropivacaine, levobupivacaine, pramoxine, lamotrigine, carbamazepine, and DEX have (0.34, 0.23, 0.21, 0.44, 0.50, and 0.40 mM, respectively). On the other hand, the IC50 values of fluoxetine, desipramine, maprotiline, and trazodone, which are 1.5, 1.6, 0.95, and ca. 1 mM, respectively, were comparable to those of lidocaine, cocaine, procaine, prilocaine, and oxymetazoline (0.74, 0.80, 2.2, and 1.8 and 1.5 mM, respectively). There was not a common chemical structure among the former (IC50 values ranging from 0.2 to 0.5 mM) or latter (IC50 values ranging from 0.7 to 2 mM) drugs, although the number of CH2 in opioids having similar structures was related to the extent of CAP inhibition (refer to Section 2.2). The antidepressants had IC50 values that were much greater than that (0.014 mM) of tetracaine. Therefore, several of the analgesic adjuvants are able to inhibit nerve AP conduction with a similar efficacy to each other.
By comparing the IC50 values of analgesic adjuvants with those of antipyretic analgesics NSAIDs, diclofenac’s IC50 value (0.94 mM) was close to the values that lidocaine, cocaine, maprotiline, and trazodone have (0.74, 0.80, 0.95, and ca. 1 mM, respectively); the IC50 values of aceclofenac, tolfenamic acid, meclofenamic acid, and flufenamic acid, which are 0.47, 0.29, 0.19, and 0.22 mM, respectively, were similar to the values that ropivacaine, levobupivacaine, pramoxine, duloxetine, amitriptyline, lamotrigine, carbamazepine, and DEX have (0.34, 0.23, 0.21, 0.23, 0.26, 0.44, 0.50, and 0.40 mM, respectively). The IC50 values of NSAIDs were less than those of procaine, prilocaine, fluoxetine, desipramine, and oxymetazoline (2.2, 1.8, 1.5, 1.6, and 1.5 mM, respectively), whereas being greater than that (0.014 mM) of tetracaine. Therefore, NSAIDs could suppress nerve AP conduction with efficiencies that are similar to those of several analgesic adjuvants. In these cases, common chemical structures were not noticed among compounds with similar IC50 values.
Aside from NSAIDs and analgesic adjuvants, narcotic analgesics, opioids, can also inhibit nerve AP conduction. Frog sciatic nerve CAP peak amplitudes were also reduced by opioids; the IC50 values of tramadol and ethylmorphine were 2.3 and 4.6 mM, respectively; morphine and codeine (each 5 mM) attenuated CAP amplitude by 15 and 30%, respectively. The magnitudes of these opioid actions were less than those of NSAIDs and analgesic adjuvants. For instance, the IC50 value (2.3 mM) of tramadol was greater than the values of lidocaine and ropivacaine, which are 0.74 and 0.34 mM, respectively, by 3.1 and 6.8 times, respectively [27]. Lidocaine has been previously reported in the frog sciatic nerve by other investigators to reduce CAP amplitudes with an IC50 value of 6.6 mM, which is greater than the tramadol value by three-fold [106]. The ratio of the IC50 value of tramadol to that of lidocaine was almost similar to that reported in the frog sciatic nerve by Katsuki et al. [27], although the IC50 values of lidocaine greatly differed between the two studies [27,106]. The extent of a contribution of nerve AP conduction inhibition to opioids-induced analgesia seems to be much less than other well-known cellular mechanisms including membrane hyperpolarization and decreased amount of L-glutamate released from nerve terminals in the spinal dorsal horn (for instance, refer to [73,74]). In conclusion, nerve AP conduction suppression may be a common cellular mechanism for NSAIDs and analgesic adjuvants, but not opioids, to cause analgesia.
The IC50 values for frog sciatic nerve CAP inhibition produced by the analgesic adjuvants were similar to those in rat sciatic nerve CAPs and were in general larger than the IC50 values for TTX-sensitive sodium channel inhibitions. This difference could be explained by several possibilities. First, CAPs are produced by not only voltage-dependent sodium channels but also potassium channels. Second, the expression of sodium channel types (Nav1.1-1.4, Nav1.6, and Nav1.7) sensitive to TTX may be different, depending on the preparations examined. Third, CAPs originate from a bundle of nerve fibers but sodium currents originate from individual cells. When the analgesic adjuvants prescribed in clinical practice act on nerve bundles, their sciatic nerve IC50 values may be a good guide to know whether the drugs suppress nerve AP conduction in vivo, as nerve conduction is due to the activation of both voltage-dependent sodium and potassium channels. Taking into consideration the experimental result that the nociceptor-specific deletion of the gene of Nav1.7 channels sensitive to TTX leads to inhibited acute and chronic pain in mice [280], sodium channels may be the primary target that analgesics and analgesic adjuvants act on.

5. Actions of Plant-Derived Compounds on Nerve AP Conduction

Many of the plant-derived compounds activate the TRP channels located in the peripheral terminals of primary-afferent Aδ-fiber and C-fiber neurons, resulting in AP production, which in turn leads to temperature sensation and nociception. For example, capsaicin, allyl isothiocyanate, and menthol open the TRPV1, TRPA1, and TRP melastatin-8 (TRPM8) channels, respectively (for example, refer to [15,281,282]). On the other hand, TRPV1, TRPA1, and TRPM8 channels are located in primary-afferent neuron central terminals in the SG of the spinal dorsal horn; the central terminal TRP channels are activated by various plant-derived compounds such as capsaicin, allyl isothiocyanate, menthol, eugenol, carvacrol, thymol, (-)-carvone, (+)-carvone, 1,4-cineole, 1,8-cineole, (±)-linalool, and geraniol (refer to [283] and reviews [284,285]). This activation is thought to be involved in the modulation of excitatory and inhibitory synaptic transmissions in SG neurons, resulting in nociceptive transmission modulation. Consistent with this idea, excitatory and inhibitory transmissions in SG neurons are affected by a variety of endogenous pain modulators (for example, refer to [17]).
As with analgesics and analgesic adjuvants, CAPs in the frog sciatic nerve were inhibited by many of the plant-derived chemicals that cause analgesia when they topically, orally, intraperitoneally, or intrathecally were applied [286,287]. Carvacrol, thymol, citronellol, bornyl acetate, citral, citronellal, geranyl acetate, and geraniol reduced frog sciatic nerve CAP peak amplitudes with the IC50 values of 0.34, 0.34, 0.35, 0.44, 0.46, 0.50, 0.51, and 0.53 mM, respectively (Table 1). Although capsaicin’s IC50 value was not able to be evaluated due to its lower solubility, capsaicin at 0.1 mM reduced CAP peak amplitudes by 36% (Table 1); this action could be attributed to at least part of the alleviation of chronic pain that is produced when capsaicin is applied to the skin ([288,289]; for example, refer to [15]). The activities of plant-derived compounds were similar to those of analgesic adjuvants and NSAIDs ([164,290,291,292]; refer to review [33]). Thus, their IC50 values were comparable to values that duloxetine, amitriptyline, aceclofenac, tolfenamic acid, meclofenamic acid, and flufenamic acid have (0.23, 0.26, 0.47 0.29, 0.19, and 0.22 mM, respectively). Furthermore, (+)-pulegone, (-)-carvone, (+)-carvone, (+)-borneol, (±)-linalool, (-)-menthone, (+)-menthone, (-)-carveol, α-terpineol, rose oxide, cinnamaldehyde, and allyl isothiocyanate attenuated frog sciatic nerve CAP peak amplitudes with IC50 values of 1.4, 1.4, 2.0, 1.5, 1.7, 1.5, 2.2, 1.3, 2.7, 2.6, 1.2, and 1.5 mM, respectively (Table 1), which were similar to those of fluoxetine and desipramine (1.5 and 1.6 mM, respectively). Linalyl acetate, eugenol, (-)-menthol, and (+)-menthol had IC50 values of 0.71, 0.81, 1.1, and 0.93 mM, respectively (Table 1); these values were close to those of diclofenac, maprotiline, and trazodone, which are 0.94, 0.95, and ca. 1.0 mM, respectively.
The cinnamaldehyde and allyl isothiocyanate activities were not affected by a non-selective TRP antagonist, ruthenium red. This result indicates that TRP channels are not involved in CAP inhibitions produced by them [291]. Capsaicin at high concentrations (0.03–0.1 mM) inhibited voltage-dependent sodium channels in rodents without TRPV1 channel activation [293,294,295]. Other plant-derived compound-induced CAP inhibitions are possibly caused by an attenuation of TTX-sensitive voltage-dependent sodium channels (for example, eugenol [296], thymol [297], carvacrol [298], and linalool [299]; refer to review [300]). The IC50 value (0.37 mM) for carvacrol to inhibit voltage-dependent sodium channels in rat DRG neurons [301] was very similar to that (0.34 mM) for the frog sciatic nerve CAP inhibition.
Some plant-derived compounds had weak inhibitory effects on frog sciatic nerve CAPs. For example, 1,8-cineole, 1,4-cineole, zingerone, guaiacol, and vanillin had IC50 values of 5.7, 7.2, 8.3, 7.7, and 9.0 mM, respectively. p-Cymene (2 mM), myrcene (5 mM), vanillylamine, and (+)-limonene (each 10 mM) reduced CAP peak amplitudes by 22, 7, 12, and 8%, respectively; vanillic acid (7 mM), p-menthane, and menthyl chloride (each 10 mM) did not affect CAPs (Table 1). Their compounds’ activities were much smaller than those of NSAIDs and analgesic adjuvants. In conclusion, some plant-derived compounds can be used instead of NSAIDs and analgesic adjuvants with respect to nerve AP conduction inhibition. Plant-derived compounds are expected to have side effects less than synthetic analgesics.
Although the above-mentioned plant-derived compounds have open chains or six-membered rings, a seven-membered ring chemical, hinokitiol (β-thujaplicin; 2-hydroxy-4-isopropylcyclohepta-2,4,6-trien-1-one, which is present in essential oils obtained from cypress trees [302]), also concentration-dependently attenuated the peak amplitudes of CAPs recorded from the frog sciatic nerve, and an IC50 value for this activity was 0.54 mM (Table 1). This value was similar to those of many other plant-derived compounds. The CAP suppression is probably mediated by certain interaction in which the carbonyl (-C=O), isopropyl (-CH(CH3)2), and hydroxyl (-OH) groups of hinokitiol are involved [303]. The -C=O group of hinokitiol helps its seven-membered ring act as a benzene ring, while the -CH(CH3)2 and -OH groups are pivotal for CAP suppression caused by hinokitiol. This idea is in favor of the observation that benzene ring compounds that have -CH(CH3)2 and -OH groups, including thymol, carvacrol, biosol (the last of which is a stereoisomer of thymol and carvacrol, and has an IC50 value of 0.58 mM), and 4-isopropylphenol (0.85 mM), can suppress frog sciatic nerve CAPs (refer to above; [290,303]). As with other plant-derived compounds, hinokitiol has a variety of actions, including suppression of apoptosis [304], the activities of anti-bacterium, anti-inflammation [305], insecticide [306], anti-fungus [307], anti-tumor [308], and cytotoxicity [309,310]. Hinokitiol, which is used as a skin medicine for inflammation reduction, may have a local anesthetic action. Applying a hinokitiol-containing oral care gel to the oral mucosa is reported to relieve oral pain in patients having oral lichen planus related to the infection of hepatitis C virus [311]. Such a pain attenuation is likely to be partly mediated by hinokitiol’s local anesthetic effect.
A CAP inhibitory action similar to hinokitiol was seen with a general anesthetic, propofol (2,6-diisopropylphenol; [312,313,314]; refer to reviews [315,316]), which has two -CH(CH3)2 groups and one -OH group attached to the benzene ring. Thus, propofol concentration-dependently reduced CAP peak amplitudes in the frog sciatic nerve and this activity had an IC50 of 0.14 mM, which is smaller than that of hinokitiol (Table 1; [29]). Consistent with this experimental result, propofol reportedly suppressed APs that were extracellularly recorded in the human and mammalian CNS [317,318]. Such an inhibitory action of propofol on nerve AP conduction may contribute to its antinociceptive effect together with a propofol-induced enhancement of GABAA receptor responses in SG neurons [319].
In clinical practice, traditional Japanese medicines (called Kampo medicines) composed of crude chemicals isolated from plants are prescribed for various purposes such as analgesia along with Western medicines in Japan [320,321,322,323,324,325]. Daikenchuto is one of most frequently prescribed Kampo medicines and is used to treat cold sensation and dysmotility in the abdomen. Frog sciatic nerve CAP amplitudes were concentration-dependently reduced by daikenchuto and by other Kampo medicines, such as rikkosan, kikyoto, rikkunshito, shakuyakukanzoto, and kakkonto. Among these medicines, daikenchuto had the greatest effect and an IC50 value for this effect was 1.1 mg/mL. Daikenchuto contains three kinds of crude medicine, such as Japanese pepper, processed ginger, and ginseng radix. The former two have been shown to suppress CAPs while the last has no effects on CAPs. Japanese pepper had an IC50 value of 0.77 mg/mL, and processed ginger at a concentration of 2 mg/mL reduced the peak amplitudes of CAPs by 31% [326]. A small part of the analgesic effect of Kampo medicine may be due to its nerve AP conduction inhibitory action.
The frog sciatic nerve CAP inhibitory action of plant-derived compounds is reversible or irreversible, depending on the type of compound. Therefore, in clinical practice, plant-derived compounds administered to peripheral nerves may or may not show tolerance in nerve AP conduction inhibition, depending on the type of compound.
Table 1. Effects of clinical analgesics, analgesic adjuvants, and plant-derived chemicals on fast-conducting frog sciatic nerve CAPs.
Table 1. Effects of clinical analgesics, analgesic adjuvants, and plant-derived chemicals on fast-conducting frog sciatic nerve CAPs.
CategoryCompoundIC50 (mM)Studied Concentration (mM)Observed AP Reduction (%)References
Acetic acid-based NSAIDsDiclofenac0.94 --[26]
Aceclofenac0.47--[26]
Indomethacin-138[26]
Acemetacin-138[26]
Etodolac-115[26]
Sulindac-10[26]
Feblinac-10[26]
Fenamic acid-based NSAIDsTolfenamic acid0.29--[26]
Meclofenamic acid0.19--[26]
Mefenamic acid-0.216[26]
Flufenamic acid0.22--[26]
Salicylic acid-based NSAIDAspirin-10[26]
Propionic acid-based NSAIDsKetoprofen-10[26]
Naproxen-10[26]
Ibuprofen-10[26]
Loxoprofen-10[26]
Flurbiprofen-10[26]
Enolic acid-based NSAIDsMeloxicam-0.50[26]
Piroxicam-10[26]
OpioidsTramadol2.3--[27]
Mono-O-desmethyl-tramadol-59[27]
Morphine-515[28]
Codeine-530[28]
Ethylmorphine4.6--[28]
Amide-type local anestheticsLidocaine0.74--[28]
Ropivacaine0.34--[27]
Prilocaine1.8--[67]
Levobupivacaine0.23--[30]
Bupivacaine-0.576[30]
Ester-type local anestheticsCocaine0.80--[28]
Procaine2.2--[164]
Benzocaine0.80--[29]
Tetracaine0.014--[32]
Other-type local anestheticPramoxine0.21--[67]
AntiepilepticsLamotrigine0.44--[30]
Carbamazepine0.50--[30]
Oxcarbazepine-0.520[30]
Phenytoin-0.115[30]
Gabapentin-100[30]
Topiramate-100[30]
Sodium valproate-100[30]
AntidepressantsDuloxetine0.23--[31]
Fluoxetine1.5--[31]
Amitriptyline0.26--[31]
Desipramine1.6--[31]
Maprotiline0.95--[31]
Trazodoneca. 1.0--[31]
Adrenoceptor agonistsAdrenaline-10[32]
Noradrenaline-10[32]
Dexmedetomidine0.40--[32]
Oxymetazoline1.5--[32]
Clonidine-2ca. 20[32]
Phenylephrine-10[32]
Isoproterenol-10[32]
Open chain or six-membered plant-derived compoundsCarvacrol0.34--[290]
Thymol0.34--[290]
Citronellol0.35--[292]
Bornyl acetate0.44--[292]
Citral0.46--[292]
Citronellal0.50--[292]
Geranyl acetate0.51--[292]
Geraniol0.53--[292]
Capsaicin-0.136[164]
(+)-Pulegone1.4--[290]
(-)-Carvone1.4--[290]
(+)-Carvone2.0--[290]
(+)-Borneol1.5--[292]
(±)-Linalool1.7--[292]
(-)-Menthone1.5--[290]
(+)-Menthone2.2--[290]
(-)-Carveol1.3--[290]
α-Terpineol2.7--[292]
Rose oxide2.6--[292]
Cinnamaldehyde1.2--[291]
Ally isothiocyanate1.5--[291]
Linalyl acetate0.71--[292]
Eugenol0.81--[164]
(-)-Menthol1.1--[290]
(+)-Menthol0.93--[290]
1,8-Cineole5.7--[290]
1,4-Cineole7.2--[290]
Zingerone8.3--[164]
Guaiacol7.7--[164]
Vanilin9.0--[164]
p-Cymene-222[292]
Myrcene-57[292]
Vanillylamine-1012[164]
(+)-Limonene-108[290]
Vanillic acid-70[164]
p-Menthane-100[290]
Menthyl chloride-100[290]
Seven-membered ring plant-derived compoundHinokitiol0.54--[303]
General
anesthetic
Propofol0.14--[29]
The IC50 value for CAP inhibition and extent of CAP amplitude reduction at each concentration are shown here.

6. Conclusions

This entry demonstrated that frog sciatic nerve CAPs are depressed by some NSAIDs, analgesic adjuvants, and plant-derived compounds that have analgesic activities with similar efficiencies, as well as opioids that have lower efficiencies than the above-mentioned categories. Although the CAPs are derived from TTX-sensitive APs generated in fast-conducting Aα-fibers, nociceptive message is transferred through slow-conducting Aδ-fibers and C-fibers [1]. Slow-conducting frog sciatic nerve CAPs could not be recorded, as Aδ-fiber CAPs were not able to be separated from Aα-fiber ones. The peak amplitude and conduction velocity of C-fiber CAP were much smaller than those of fast-conducting Aα ones [25]. Therefore, it was not possible to investigate how slow-conducting CAPs are affected by the antinociceptive compounds. In order to know details regarding the differences in the magnitude of nerve AP conduction suppression among a variety of analgesic compounds, it would be required to investigate how slow-conducting CAPs are affected by the compounds.
In nerves other than the sciatic nerve in frogs, antinociceptive drugs have reportedly suppressed both A-fiber and C-fiber CAPs. For instance, in the vagus nerve in rabbits, fentanyl and sufentanil inhibited C-fiber CAPs with magnitudes less than those observed in A-fiber ones [85], and lidocaine suppressed both myelinated A-fiber and unmyelinated C-fiber nerve AP conduction ([327]; lidocaine inhibited rat A-fiber CAPs more effectively than C-fiber CAPs [328]). Clonidine produced both Aα-fiber and C-fiber CAP amplitude reductions (refer to Section 3.4). Furthermore, many researchers have demonstrated a suppression by NSAIDs [56,63,329], lidocaine, and α2-adrenoceptor agonists [145] of TTX-resistant voltage-dependent sodium channels, which may be involved in slow-conducting CAP production. Knocking down Nav1.8 channels resistant to TTX reportedly resulted in suppressed pain associated with neuropathy or inflammation in rats [330].
Because the analgesic and analgesic adjuvant concentrations needed for CAP inhibition are higher than clinically relevant ones (as mentioned above), nerve AP conduction inhibition may only occur when the drugs are administered near nerve fibers or are accumulated in the nervous system. If such a suppression is produced in Aα-fibers that innervate skeletal muscle, this would produce unwanted secondary effects, such as muscle paralysis. Thus, the drug will have to be used at the lowest concentration while it is effective. Primary-afferent C-fibers and Aδ-fibers have diameters smaller than those of Aα-fibers. Therefore, if the drugs act on voltage-dependent sodium channels from the cytoplasm side, nerve AP conduction would be inhibited in C-fibers earlier than in Aα-fibers because the surface-to-volume ratio differs between the fibers. Because at least part of the analgesic action of most agents relies on the suppression of nerve AP conduction mediated by voltage-dependent TTX-sensitive and TTX-resistant sodium channels, the careful use of these compound aids in the study of nociception as well as clinical pain relief.

Funding

This research received no external funding.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Fields, H.L. Pain; McGraw-Hill: New York, NY, USA, 1987. [Google Scholar]
  2. Willis, W.D., Jr.; Coggeshall, R.E. Sensory Mechanisms of the Spinal Cord, 2nd ed.; Plenum: New York, NY, USA, 1991. [Google Scholar]
  3. Todd, A.J. Neuronal circuitry for pain processing in the dorsal horn. Nat. Rev. Neurosci. 2010, 11, 823–836. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Merighi, A. The histology, physiology, neurochemistry and circuitry of the substantia gelatinosa Rolandi (lamina II) in mammalian spinal cord. Prog. Neurobiol. 2018, 169, 91–134. [Google Scholar] [PubMed]
  5. Merskey, H. Clarifying definition of neuropathic pain. Pain 2002, 96, 408–409. [Google Scholar] [CrossRef]
  6. Amir, R.; Argoff, C.E.; Bennett, G.J.; Cummins, T.R.; Durieux, M.E.; Gerner, P.; Gold, M.S.; Porreca, F.; Strichartz, G.R. The role of sodium channels in chronic inflammatory and neuropathic pain. J. Pain 2006, 7, S1–S29. [Google Scholar] [CrossRef]
  7. Finnerup, N.B.; Sindrup, S.H.; Jensen, T.S. The evidence for pharmacological treatment of neuropathic pain. Pain 2010, 150, 573–581. [Google Scholar] [CrossRef] [PubMed]
  8. Jensen, T.S. Anticonvulsants in neuropathic pain: Rationale and clinical evidence. Eur. J. Pain 2002, 6, 61–68. [Google Scholar] [CrossRef] [PubMed]
  9. Kamibayashi, T.; Maze, M. Clinical uses of α2-adrenergic agonists. Anesthesiology 2000, 93, 1345–1349. [Google Scholar] [CrossRef] [PubMed]
  10. Lynch, M.E. Antidepressants as analgesics: A review of randomized controlled trials. J. Psychiatry Neurosci. 2001, 26, 30–36. [Google Scholar] [PubMed]
  11. Sindrup, S.H.; Otto, M.; Finnerup, N.B.; Jensen, T.S. Antidepressants in the treatment of neuropathic pain. Basic Clin. Pharmacol. Toxicol. 2005, 96, 399–409. [Google Scholar] [CrossRef] [PubMed]
  12. Theile, J.W.; Cummins, T.R. Recent developments regarding voltage-gated sodium channel blockers for the treatment of inherited and acquired neuropathic pain syndromes. Front. Pharmacol. 2011, 2, 54. [Google Scholar] [CrossRef] [PubMed]
  13. Waszkielewicz, A.M.; Gunia, A.; Słoczyńska, K.; Marona, H. Evaluation of anticonvulsants for possible use in neuropathic pain. Curr. Med. Chem. 2011, 18, 4344–4358. [Google Scholar] [CrossRef] [PubMed]
  14. Fakhri, S.; Abbaszadeh, F.; Jorjani, M. On the therapeutic targets and pharmacological treatments for pain relief following spinal cord injury: A mechanistic review. Biomed. Pharmacother. 2021, 139, 111563. [Google Scholar] [PubMed]
  15. Kocot-Kępska, M.; Zajączkowska, R.; Mika, J.; Kopsky, D.J.; Wordliczek, J.; Dobrogowski, J.; Przeklasa-Muszyńska, A. Topical treatments and their molecular/cellular mechanisms in patients with peripheral neuropathic pain—Narrative review. Pharmaceutics 2021, 13, 450. [Google Scholar] [CrossRef] [PubMed]
  16. Fürst, S. Transmitters involved in antinociception in the spinal cord. Brain Res. Bull. 1999, 48, 129–141. [Google Scholar] [CrossRef] [PubMed]
  17. Kumamoto, E. Cellular mechanisms for antinociception produced by oxytocin and orexins in the rat spinal lamina II—Comparison with those of other endogenous pain modulators. Pharmaceuticals 2019, 12, 136. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Zeilhofer, H.U.; Wildner, H.; Yévenes, G.E. Fast synaptic inhibition in spinal sensory processing and pain control. Physiol. Rev. 2012, 92, 193–235. [Google Scholar] [CrossRef]
  19. Gouveia, D.N.; Pina, L.T.S.; Rabelo, T.K.; da Rocha Santos, W.B.; Quintans, J.S.S.; Guimarães, A.G. Monoterpenes as perspective to chronic pain management: A systematic review. Curr. Drug Targets 2018, 19, 960–972. [Google Scholar] [CrossRef]
  20. Wang, Z.-J.; Heinbockel, T. Essential oils and their constituents targeting the GABAergic system and sodium channels as treatment of neurological diseases. Molecules 2018, 23, 1061. [Google Scholar] [CrossRef] [Green Version]
  21. Gouveia, D.N.; Guimarães, A.G.; da Rocha Santos, W.B.; Quintans-Júnior, L.J. Natural products as a perspective for cancer pain management: A systematic review. Phytomedicine 2019, 58, 152766. [Google Scholar] [CrossRef]
  22. Kiernan, M.C.; Bostock, H.; Park, S.B.; Kaji, R.; Krarup, C.; Krishnan, A.V.; Kuwabara, S.; Lin, C.S.; Misawa, S.; Moldovan, M.; et al. Measurement of axonal excitability: Consensus guidelines. Clin. Neurophysiol. 2020, 131, 308–323. [Google Scholar] [CrossRef]
  23. Levitan, I.B.; Karczmarek, L.K. The Neuron, 3rd ed.; Oxford University Press: New York, NY, USA, 2002. [Google Scholar]
  24. Kumamoto, E.; Mizuta, K.; Fujita, T. Peripheral nervous system in the frog as a tool to examine the regulation of the transmission of neuronal information. In Frogs: Biology, Ecology and Uses; Murray, J.L., Ed.; Nova Science Publishers, Inc.: New York, NY, USA, 2012; pp. 89–106. [Google Scholar]
  25. Kobayashi, J.; Ohta, M.; Terada, Y. C fiber generates a slow Na+ spike in the frog sciatic nerve. Neurosci. Lett. 1993, 162, 93–96. [Google Scholar] [CrossRef] [PubMed]
  26. Suzuki, R.; Fujita, T.; Mizuta, K.; Kumamoto, E. Inhibition by non-steroidal anti-inflammatory drugs of compound action potentials in frog sciatic nerve fibers. Biomed. Pharmacother. 2018, 103, 326–335. [Google Scholar] [CrossRef] [PubMed]
  27. Katsuki, R.; Fujita, T.; Koga, A.; Liu, T.; Nakatsuka, T.; Nakashima, M.; Kumamoto, E. Tramadol, but not its major metabolite (mono-O-demethyl tramadol) depresses compound action potentials in frog sciatic nerves. Br. J. Pharmacol. 2006, 149, 319–327. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Mizuta, K.; Fujita, T.; Nakatsuka, T.; Kumamoto, E. Inhibitory effects of opioids on compound action potentials in frog sciatic nerves and their chemical structures. Life Sci. 2008, 83, 198–207. [Google Scholar] [CrossRef] [PubMed]
  29. Magori, N.; Fujita, T.; Mizuta, K.; Kumamoto, E. Inhibition by general anesthetic propofol of compound action potentials in the frog sciatic nerve and its chemical structure. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2019, 392, 359–369. [Google Scholar] [CrossRef] [PubMed]
  30. Uemura, Y.; Fujita, T.; Ohtsubo, S.; Hirakawa, N.; Sakaguchi, Y.; Kumamoto, E. Effects of various antiepileptics used to alleviate neuropathic pain on compound action potential in frog sciatic nerves: Comparison with those of local anesthetics. Biomed. Res. Int. 2014, 2014, 540238. [Google Scholar] [CrossRef] [PubMed]
  31. Hirao, R.; Fujita, T.; Sakai, A.; Kumamoto, E. Compound action potential inhibition produced by various antidepressants in the frog sciatic nerve. Eur. J. Pharmacol. 2018, 819, 122–128. [Google Scholar] [CrossRef] [PubMed]
  32. Kosugi, T.; Mizuta, K.; Fujita, T.; Nakashima, M.; Kumamoto, E. High concentrations of dexmedetomidine inhibit compound action potentials in frog sciatic nerves without α2 adrenoceptor activation. Br. J. Pharmacol. 2010, 160, 1662–1676. [Google Scholar] [CrossRef] [Green Version]
  33. Kumamoto, E. Effects of plant-derived compounds on excitatory synaptic transmission and nerve conduction in the nervous system—Involvement in pain modulation. Curr. Top. Phytochem. 2018, 14, 45–70. [Google Scholar]
  34. Ferreira, S.H. Prostaglandins, aspirin-like drugs and analgesia. Nat. New Biol. 1972, 240, 200–203. [Google Scholar] [CrossRef]
  35. Takayama, K.; Hirose, A.; Suda, I.; Miyazaki, A.; Oguchi, M.; Onotogi, M.; Fotopoulos, G. Comparison of the anti-inflammatory and analgesic effects in rats of diclofenac-sodium, felbinac and indomethacin patches. Int. J. Biomed. Sci. 2011, 7, 222–229. [Google Scholar] [PubMed]
  36. Vane, J.R. Introduction: Mechanism of action of NSAIDs. Br. J. Rheumatol. 1996, 35 (Suppl. S1), 1–3. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Simmons, D.L.; Botting, R.M.; Hla, T. Cyclooxygenase isozymes: The biology of prostaglandin synthesis and inhibition. Pharmacol. Rev. 2004, 56, 387–437. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Grosser, T.; Smyth, E.; FitzGerald, G.A. Anti-inflammatory, antipyretic, and analgesic agents; pharmacotherapy of gout. In Goodman & Gilman’s the Pharmacological Basis of Therapeutics, 12th ed.; Brunton, L.L., Chabner, B.A., Knollmann, B.C., Eds.; McGraw-Hill, Medical Publishing Division: New York, NY, USA, 2011; pp. 959–1004. [Google Scholar]
  39. Kaduševičius, E. Novel applications of NSAIDs: Insight and future perspectives in cardiovascular, neurodegenerative, diabetes and cancer disease therapy. Int. J. Mol. Sci. 2021, 22, 6637. [Google Scholar] [CrossRef] [PubMed]
  40. Garg, P.; Sanguinetti, M.C. Structure-activity relationship of fenamates as Slo2.1 channel activators. Mol. Pharmacol. 2012, 82, 795–802. [Google Scholar] [CrossRef] [Green Version]
  41. Ortiz, M.I.; Torres-López, J.E.; Castañeda-Hernández, G.; Rosas, R.; Vidal-Cantú, G.C.; Granados-Soto, V. Pharmacological evidence for the activation of K+ channels by diclofenac. Eur. J. Pharmacol. 2002, 438, 85–91. [Google Scholar] [CrossRef]
  42. Ortiz, M.I.; Castañeda-Hernández, G.; Granados-Soto, V. Pharmacological evidence for the activation of Ca2+-activated K+ channels by meloxicam in the formalin test. Pharmacol. Biochem. Behav. 2005, 81, 725–731. [Google Scholar] [CrossRef]
  43. Ortiz, M.I.; Granados-Soto, V.; Castañeda-Hernández, G. The NO-cGMP-K+ channel pathway participates in the antinociceptive effect of diclofenac, but not of indomethacin. Pharmacol. Biochem. Behav. 2003, 76, 187–195. [Google Scholar] [CrossRef]
  44. Peretz, A.; Degani, N.; Nachman, R.; Uziyel, Y.; Gibor, G.; Shabat, D.; Attali, B. Meclofenamic acid and diclofenac, novel templates of KCNQ2/Q3 potassium channel openers, depress cortical neuron activity and exhibit anticonvulsant properties. Mol. Pharmacol. 2005, 67, 1053–1066. [Google Scholar] [CrossRef] [Green Version]
  45. Gwanyanya, A.; Macianskiene, R.; Mubagwa, K. Insights into the effects of diclofenac and other non-steroidal anti-inflammatory agents on ion channels. J. Pharm. Pharmacol. 2012, 64, 1359–1375. [Google Scholar] [CrossRef]
  46. Guinamard, R.; Simard, C.; Del Negro, C. Flufenamic acid as an ion channel modulator. Pharmacol. Ther. 2013, 138, 272–284. [Google Scholar] [CrossRef]
  47. Voilley, N.; de Weille, J.; Mamet, J.; Lazdunski, M. Nonsteroid anti-inflammatory drugs inhibit both the activity and the inflammation-induced expression of acid-sensing ion channels in nociceptors. J. Neurosci. 2001, 21, 8026–8033. [Google Scholar] [CrossRef] [PubMed]
  48. Inoue, N.; Ito, S.; Nogawa, M.; Tajima, K.; Kyoi, T. Etodolac blocks the allyl isothiocyanate-induced response in mouse sensory neurons by selective TRPA1 activation. Pharmacology 2012, 90, 47–54. [Google Scholar] [CrossRef] [PubMed]
  49. Suzuki, H.; Sasaki, E.; Nakagawa, A.; Muraki, Y.; Hatano, N.; Muraki, K. Diclofenac, a nonsteroidal anti-inflammatory drug, is an antagonist of human TRPM3 isoforms. Pharmacol. Res. Perspect. 2016, 4, e00232. [Google Scholar] [CrossRef]
  50. Papworth, J.; Colville-Nash, P.; Alam, C.; Seed, M.; Willoughby, D. The depletion of substance P by diclofenac in the mouse. Eur. J. Pharmacol. 1997, 325, R1–R2. [Google Scholar] [CrossRef] [PubMed]
  51. Silva, L.C.R.; Castor, M.G.M.; Souza, T.C.; Duarte, I.D.G.; Romero, T.R.L. NSAIDs induce peripheral antinociception by interaction with the adrenergic system. Life Sci. 2015, 130, 7–11. [Google Scholar] [CrossRef]
  52. Vazquez, E.; Hernandez, N.; Escobar, W.; Vanegas, H. Antinociception induced by intravenous dipyrone (metamizol) upon dorsal horn neurons: Involvement of endogenous opioids at the periaqueductal gray matter, the nucleus raphe magnus, and the spinal cord in rats. Brain Res. 2005, 1048, 211–217. [Google Scholar] [CrossRef]
  53. Silva, L.C.R.; Castor, M.G.M.; Navarro, L.C.; Romero, T.R.L.; Duarte, I.D.G. κ-Opioid receptor participates of NSAIDs peripheral antinociception. Neurosci. Lett. 2016, 622, 6–9. [Google Scholar] [CrossRef]
  54. Fowler, C.J. NSAIDs: eNdocannabinoid stimulating anti-inflammatory drugs? Trends Pharmacol. Sci. 2012, 33, 468–473. [Google Scholar] [CrossRef]
  55. McCormack, K.; Brune, K. Dissociation between the antinociceptive and anti-inflammatory effects of the nonsteroidal anti-inflammatory drugs. A survey of their analgesic efficacy. Drugs 1991, 41, 533–547. [Google Scholar] [CrossRef]
  56. Lee, H.M.; Kim, H.I.; Shin, Y.K.; Lee, C.S.; Park, M.; Song, J.-H. Diclofenac inhibition of sodium currents in rat dorsal root ganglion neurons. Brain Res. 2003, 992, 120–127. [Google Scholar] [CrossRef] [PubMed]
  57. Acosta, M.C.; Luna, C.; Graff, G.; Meseguer, V.M.; Viana, F.; Gallar, J.; Belmonte, C. Comparative effects of the nonsteroidal anti-inflammatory drug nepafenac on corneal sensory nerve fibers responding to chemical irritation. Investig. Ophthalmol. Vis. Sci. 2007, 48, 182–188. [Google Scholar] [CrossRef] [Green Version]
  58. Fei, X.-W.; Liu, L.-Y.; Xu, J.-G.; Zhang, Z.-H.; Mei, Y.-A. The non-steroidal anti-inflammatory drug, diclofenac, inhibits Na+ current in rat myoblasts. Biochem. Biophys. Res. Commun. 2006, 346, 1275–1283. [Google Scholar] [CrossRef] [PubMed]
  59. Yarishkin, O.V.; Hwang, E.M.; Kim, D.; Yoo, J.C.; Kang, S.S.; Kim, D.R.; Shin, J.-H.-J.; Chung, H.-J.; Jeong, H.-S.; Kang, D.; et al. Diclofenac, a non-steroidal anti-inflammatory drug, inhibits L-type Ca2+ channels in neonatal rat ventricular cardiomyocytes. Korean J. Physiol. Pharmacol. 2009, 13, 437–442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Kuo, C.-C.; Huang, R.-C.; Lou, B.-S. Inhibition of Na+ current by diphenhydramine and other diphenyl compounds: Molecular determinants of selective binding to the inactivated channels. Mol. Pharmacol. 2000, 57, 135–143. [Google Scholar]
  61. Yang, Y.-C.; Kuo, C.-C. An inactivation stabilizer of the Na+ channel acts as an opportunistic pore blocker modulated by external Na+. J. Gen. Physiol. 2005, 125, 465–481. [Google Scholar] [CrossRef] [Green Version]
  62. Yau, H.-J.; Baranauskas, G.; Martina, M. Flufenamic acid decreases neuronal excitability through modulation of voltage-gated sodium channel gating. J. Physiol. 2010, 588, 3869–3882. [Google Scholar] [CrossRef]
  63. Nakamura, M.; Jang, I.-S. pH-dependent inhibition of tetrodotoxin-resistant Na+ channels by diclofenac in rat nociceptive neurons. Prog. Neuropsychopharmacol. Biol. Psychiatry 2016, 64, 35–43. [Google Scholar] [CrossRef]
  64. Sun, J.-F.; Xu, Y.-J.; Kong, X.-H.; Su, Y.; Wang, Z.-Y. Fenamates inhibit human sodium channel Nav1.7 and Nav1.8. Neurosci. Lett. 2019, 696, 67–73. [Google Scholar] [CrossRef]
  65. Chen, X.; Gallar, J.; Belmonte, C. Reduction by antiinflammatory drugs of the response of corneal sensory nerve fibers to chemical irritation. Investig. Ophthalmol. Vis. Sci. 1997, 38, 1944–1953. [Google Scholar]
  66. Kumamoto, E. Inhibition of fast nerve conduction produced by analgesics and analgesic adjuvants—Possible involvement in pain alleviation. Pharmaceuticals 2020, 13, 62. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Mizuta, K.; Fujita, T.; Yamagata, H.; Kumamoto, E. Bisphenol A inhibits compound action potentials in the frog sciatic nerve in a manner independent of estrogen receptors. Biochem. Biophys. Rep. 2017, 10, 145–151. [Google Scholar] [CrossRef] [PubMed]
  68. Gil-Flores, M.; Ortiz, M.I.; Castañeda-Hernández, G.; Chávez-Piña, A.E. Acemetacin antinociceptive mechanism is not related to NO or K+ channel pathways. Methods Find. Exp. Clin. Pharmacol. 2010, 32, 101–105. [Google Scholar] [CrossRef] [PubMed]
  69. Gögelein, H.; Dahlem, D.; Englert, H.C.; Lang, H.J. Flufenamic acid, mefenamic acid and niflumic acid inhibit single nonselective cation channels in the rat exocrine pancreas. FEBS Lett. 1990, 268, 79–82. [Google Scholar] [CrossRef] [Green Version]
  70. Hu, H.; Tian, J.; Zhu, Y.; Wang, C.; Xiao, R.; Herz, J.M.; Wood, J.D.; Zhu, M.X. Activation of TRPA1 channels by fenamate nonsteroidal anti-inflammatory drugs. Pflüg. Arch. 2010, 459, 579–592. [Google Scholar] [CrossRef] [Green Version]
  71. Tatematsu, Y.; Hayashi, H.; Taguchi, R.; Fujita, H.; Yamamoto, A.; Ohkura, K. Effect of N-phenylanthranilic acid scaffold nonsteroidal anti-inflammatory drugs on the mitochondrial permeability transition. Biol. Pharm. Bull. 2016, 39, 278–284. [Google Scholar] [CrossRef] [Green Version]
  72. Glass, J.S.; Hardy, C.L.; Meeks, N.M.; Carroll, B.T. Acute pain management in dermatology: Risk assessment and treatment. J. Am. Acad. Dermatol. 2015, 73, 543–560. [Google Scholar] [CrossRef]
  73. Fujita, T.; Kumamoto, E. Inhibition by endomorphin-1 and endomorphin-2 of excitatory transmission in adult rat substantia gelatinosa neurons. Neuroscience 2006, 139, 1095–1105. [Google Scholar] [CrossRef]
  74. Kohno, T.; Kumamoto, E.; Higashi, H.; Shimoji, K.; Yoshimura, M. Actions of opioids on excitatory and inhibitory transmission in substantia gelatinosa of adult rat spinal cord. J. Physiol. 1999, 518, 803–813. [Google Scholar] [CrossRef]
  75. Yoshimura, M.; North, R.A. Substantia gelatinosa neurones hyperpolarized in vitro by enkephalin. Nature 1983, 305, 529–530. [Google Scholar] [CrossRef]
  76. North, R.A. Opioid actions on membrane ion channels. In Handbook of Experimental Pharmacology; Herz, A., Ed.; Springer: Berlin/Heidelberg, Germany, 1993; Volume 104, pp. 773–797. [Google Scholar]
  77. Yaksh, T.L. Pharmacology and mechanisms of opioid analgesic activity. Acta Anaesthesiol. Scand. 1997, 41, 94–111. [Google Scholar] [CrossRef] [PubMed]
  78. Smith, T.W.; Buchan, P.; Parsons, D.N.; Wilkinson, S. Peripheral antinociceptive effects of N-methyl morphine. Life Sci. 1982, 31, 1205–1208. [Google Scholar] [CrossRef] [PubMed]
  79. Stein, C.; Comisel, K.; Haimerl, E.; Yassouridis, A.; Lehrberger, K.; Herz, A.; Peter, K. Analgesic effect of intraarticular morphine after arthroscopic knee surgery. N. Engl. J. Med. 1991, 325, 1123–1126. [Google Scholar] [CrossRef] [PubMed]
  80. Shannon, H.E.; Lutz, E.A. Comparison of the peripheral and central effects of the opioid agonists loperamide and morphine in the formalin test in rats. Neuropharmacology 2002, 42, 253–261. [Google Scholar] [CrossRef]
  81. Wenk, H.N.; Brederson, J.-D.; Honda, C.N. Morphine directly inhibits nociceptors in inflamed skin. J. Neurophysiol. 2006, 95, 2083–2097. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Labuz, D.; Mousa, S.A.; Schäfer, M.; Stein, C.; Machelska, H. Relative contribution of peripheral versus central opioid receptors to antinociception. Brain Res. 2007, 1160, 30–38. [Google Scholar] [CrossRef]
  83. Stein, C.; Schäfer, M.; Machelska, H. Attacking pain at its source: New perspectives on opioids. Nat. Med. 2003, 9, 1003–1008. [Google Scholar] [CrossRef]
  84. Yuge, O.; Matsumoto, M.; Kitahata, L.M.; Collins, J.G.; Senami, M. Direct opioid application to peripheral nerves does not alter compound action potentials. Anesth. Analg. 1985, 64, 667–671. [Google Scholar] [CrossRef]
  85. Gissen, A.J.; Gugino, L.D.; Datta, S.; Miller, J.; Covino, B.G. Effects of fentanyl and sufentanil on peripheral mammalian nerves. Anesth. Analg. 1987, 66, 1272–1276. [Google Scholar] [CrossRef]
  86. Jaffe, R.A.; Rowe, M.A. A comparison of the local anesthetic effects of meperidine, fentanyl, and sufentanil on dorsal root axons. Anesth. Analg. 1996, 83, 776–781. [Google Scholar] [CrossRef]
  87. Jurna, I.; Grossmann, W. The effect of morphine on mammalian nerve fibres. Eur. J. Pharmacol. 1977, 44, 339–348. [Google Scholar] [CrossRef] [PubMed]
  88. Coggeshall, R.E.; Zhou, S.; Carlton, S.M. Opioid receptors on peripheral sensory axons. Brain Res. 1997, 764, 126–132. [Google Scholar] [CrossRef] [PubMed]
  89. Fields, H.L.; Emson, P.C.; Leigh, B.K.; Gilbert, R.F.T.; Iversen, L.L. Multiple opiate receptor sites on primary afferent fibres. Nature 1980, 284, 351–353. [Google Scholar] [CrossRef]
  90. Wenk, H.N.; Honda, C.N. Immunohistochemical localization of delta opioid receptors in peripheral tissues. J. Comp. Neurol. 1999, 408, 567–579. [Google Scholar] [CrossRef]
  91. Klotz, U. Tramadol—The impact of its pharmacokinetic and pharmacodynamic properties on the clinical management of pain. Arzneimittelforschung 2003, 53, 681–687. [Google Scholar] [CrossRef]
  92. Lintz, W.; Erlacin, S.; Frankus, E.; Uragg, H. Metabolismus von Tramadol bei Mensch und Tier. Arzneimittelforschung 1981, 31, 1932–1943. [Google Scholar]
  93. Hennies, H.-H.; Friderichs, E.; Schneider, J. Receptor binding, analgesic and antitussive potency of tramadol and other selected opioids. Arzneimittelforschung 1988, 38, 877–880. [Google Scholar]
  94. Raffa, R.B.; Friderichs, E.; Reimann, W.; Shank, R.P.; Codd, E.E.; Vaught, J.L. Opioid and nonopioid components independently contribute to the mechanism of action of tramadol, an ‘atypical’ opioid analgesic. J. Pharmacol. Exp. Ther. 1992, 260, 275–285. [Google Scholar]
  95. Koga, A.; Fujita, T.; Totoki, T.; Kumamoto, E. Tramadol produces outward currents by activating µ-opioid receptors in adult rat substantia gelatinosa neurones. Br. J. Pharmacol. 2005, 145, 602–607. [Google Scholar] [CrossRef] [Green Version]
  96. Koga, A.; Fujita, T.; Piao, L.-H.; Nakatsuka, T.; Kumamoto, E. Inhibition by O-desmethyltramadol of glutamatergic excitatory transmission in adult rat spinal substantia gelatinosa neurons. Mol. Pain 2019, 15, 1744806918824243. [Google Scholar] [CrossRef] [Green Version]
  97. Yamasaki, H.; Funai, Y.; Funao, T.; Mori, T.; Nishikawa, K. Effects of tramadol on substantia gelatinosa neurons in the rat spinal cord: An in vivo patch-clamp analysis. PLoS ONE 2015, 10, e0125147. [Google Scholar] [CrossRef] [PubMed]
  98. Altunkaya, H.; Ozer, Y.; Kargi, E.; Babuccu, O. Comparison of local anaesthetic effects of tramadol with prilocaine for minor surgical procedures. Br. J. Anaesth. 2003, 90, 320–322. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Altunkaya, H.; Ozer, Y.; Kargi, E.; Ozkocak, I.; Hosnuter, M.; Demirel, C.B.; Babuccu, O. The postoperative analgesic effect of tramadol when used as subcutaneous local anesthetic. Anesth. Analg. 2004, 99, 1461–1464. [Google Scholar] [CrossRef] [PubMed]
  100. Pang, W.-W.; Mok, M.S.; Chang, D.-P.; Huang, M.-H. Local anesthetic effect of tramadol, metoclopramide, and lidocaine following intradermal injection. Reg. Anesth. Pain Med. 1998, 23, 580–583. [Google Scholar] [CrossRef] [PubMed]
  101. Le Roux, P.J.; Coetzee, J.F. Tramadol today. Curr. Opin. Anaesth. 2000, 13, 457–461. [Google Scholar] [CrossRef]
  102. Tsai, Y.-C.; Chang, P.-J.; Jou, I.-M. Direct tramadol application on sciatic nerve inhibits spinal somatosensory evoked potentials in rats. Anesth. Analg. 2001, 92, 1547–1551. [Google Scholar] [CrossRef]
  103. Shin, H.-W.; Ju, B.-J.; Jang, Y.-K.; You, H.-S.; Kang, H.; Park, J.-Y. Effect of tramadol as an adjuvant to local anesthetics for brachial plexus block: A systematic review and meta-analysis. PLoS ONE 2017, 12, e0184649. [Google Scholar] [CrossRef] [Green Version]
  104. Mert, T.; Gunes, Y.; Guven, M.; Gunay, I.; Ozcengiz, D. Comparison of nerve conduction blocks by an opioid and a local anesthetic. Eur. J. Pharmacol. 2002, 439, 77–81. [Google Scholar] [CrossRef]
  105. Güven, M.; Mert, T.; Günay, I. Effects of tramadol on nerve action potentials in rat: Comparisons with benzocaine and lidocaine. Int. J. Neurosci. 2005, 115, 339–349. [Google Scholar] [CrossRef]
  106. Mert, T.; Gunes, Y.; Guven, M.; Gunay, I.; Gocmen, C. Differential effects of lidocaine and tramadol on modified nerve impulse by 4-aminopyridine in rats. Pharmacology 2003, 69, 68–73. [Google Scholar] [CrossRef]
  107. Gillen, C.; Haurand, M.; Kobelt, D.J.; Wnendt, S. Affinity, potency and efficacy of tramadol and its metabolites at the cloned human µ-opioid receptor. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2000, 362, 116–121. [Google Scholar] [CrossRef] [PubMed]
  108. Driessen, B.; Reimann, W. Interaction of the central analgesic, tramadol, with the uptake and release of 5-hydroxytryptamine in the rat brain in vitro. Br. J. Pharmacol. 1992, 105, 147–151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Driessen, B.; Reimann, W.; Giertz, H. Effects of the central analgesic tramadol on the uptake and release of noradrenaline and dopamine in vitro. Br. J. Pharmacol. 1993, 108, 806–811. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Leffler, A.; Frank, G.; Kistner, K.; Niedermirtl, F.; Koppert, W.; Reeh, P.W.; Nau, C. Local anesthetic-like inhibition of voltage-gated Na+ channels by the partial μ-opioid receptor agonist buprenorphine. Anesthesiology 2012, 116, 1335–1346. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Haeseler, G.; Foadi, N.; Ahrens, J.; Dengler, R.; Hecker, H.; Leuwer, M. Tramadol, fentanyl and sufentanil but not morphine block voltage-operated sodium channels. Pain 2006, 126, 234–244. [Google Scholar] [CrossRef] [PubMed]
  112. Tsai, T.-Y.; Tsai, Y.-C.; Wu, S.-N.; Liu, Y.-C. Tramadol-induced blockade of delayed rectifier potassium current in NG108-15 neuronal cells. Eur. J. Pain 2006, 10, 597–601. [Google Scholar] [CrossRef]
  113. Grond, S.; Meuser, T.; Uragg, H.; Stahlberg, H.J.; Lehmann, K.A. Serum concentrations of tramadol enantiomers during patient-controlled analgesia. Br. J. Clin. Pharmacol. 1999, 48, 254–257. [Google Scholar] [CrossRef] [Green Version]
  114. Brodin, P.; Skoglund, L.A. Dose-response inhibition of rat compound nerve action potential by dextropropoxyphene and codeine compared to morphine and cocaine in vitro. Gen. Pharmacol. 1990, 21, 551–553. [Google Scholar] [CrossRef]
  115. Hunter, E.G.; Frank, G.B. An opiate receptor on frog sciatic nerve axons. Can. J. Physiol. Pharmacol. 1979, 57, 1171–1174. [Google Scholar] [CrossRef]
  116. Kumamoto, E.; Mizuta, K.; Fujita, T. Opioid actions in primary-afferent fibers—Involvement in analgesia and anesthesia. Pharmaceuticals 2011, 4, 343–365. [Google Scholar] [CrossRef] [Green Version]
  117. Bräu, M.E.; Nau, C.; Hempelmann, G.; Vogel, W. Local anesthetics potently block a potential insensitive potassium channel in myelinated nerve. J. Gen. Physiol. 1995, 105, 485–505. [Google Scholar] [CrossRef] [PubMed]
  118. Bräu, M.E.; Vogel, W.; Hempelmann, G. Fundamental properties of local anesthetics: Half-maximal blocking concentrations for tonic block of Na+ and K+ channels in peripheral nerve. Anesth. Analg. 1998, 87, 885–889. [Google Scholar] [PubMed]
  119. Tokuno, H.A.; Bradberry, C.W.; Everill, B.; Agulian, S.K.; Wilkes, S.; Baldwin, R.M.; Tamagnan, G.D.; Kocsis, J.D. Local anesthetic effects of cocaethylene and isopropylcocaine on rat peripheral nerves. Brain Res. 2004, 996, 159–167. [Google Scholar] [CrossRef] [PubMed]
  120. Chen, Z.R.; Irvine, R.J.; Somogyi, A.A.; Bochner, F. Mu receptor binding of some commonly used opioids and their metabolites. Life Sci. 1991, 48, 2165–2171. [Google Scholar] [CrossRef]
  121. Mizuta, K.; Fujita, T.; Kumamoto, E. Inhibition by morphine and its analogs of action potentials in adult rat dorsal root ganglion neurons. J. Neurosci. Res. 2012, 90, 1830–1841. [Google Scholar] [CrossRef]
  122. Staiman, A.; Seeman, P. The impulse-blocking concentrations of anesthetics, alcohols, anticonvulsants, barbiturates, and narcotics on phrenic and sciatic nerves. Can. J. Physiol. Pharmacol. 1974, 52, 535–550. [Google Scholar] [CrossRef]
  123. Scholz, A. Mechanisms of (local) anaesthetics on voltage-gated sodium and other ion channels. Br. J. Anaesth. 2002, 89, 52–61. [Google Scholar] [CrossRef] [Green Version]
  124. Hu, S.; Rubly, N. Effects of morphine on ionic currents in frog node of Ranvier. Eur. J. Pharmacol. 1983, 95, 185–192. [Google Scholar] [CrossRef]
  125. Frazier, D.T.; Murayama, K.; Abbott, N.J.; Narahashi, T. Effects of morphine on internally perfused squid giant axons. Proc. Soc. Exp. Biol. Med. 1972, 139, 434–438. [Google Scholar] [CrossRef]
  126. Wagner, L.E., II; Eaton, M.; Sabnis, S.S.; Gingrich, K.J. Meperidine and lidocaine block of recombinant voltage-dependent Na+ channels: Evidence that meperidine is a local anesthetic. Anesthesiology 1999, 91, 1481–1490. [Google Scholar] [CrossRef] [Green Version]
  127. Gutstein, H.B.; Akil, H. Opioid analgesics. In Goodman & Gilman’s the Pharmacological Basis of Therapeutics, 11th ed.; Brunton, L.L., Lazo, J.S., Parker, K.L., Eds.; McGraw-Hill, Medical Publishing Division: New York, NY, USA, 2006; pp. 547–590. [Google Scholar]
  128. Viel, E.J.; Eledjam, J.J.; De La Coussaye, J.E.; D’Athis, F. Brachial plexus block with opioids for postoperative pain relief: Comparison between buprenorphine and morphine. Reg. Anesth. 1989, 14, 274–278. [Google Scholar] [PubMed]
  129. King, M.; Su, W.; Chang, A.; Zuckerman, A.; Pasternak, G.W. Transport of opioids from the brain to the periphery by P-glycoprotein: Peripheral actions of central drugs. Nat. Neurosci. 2001, 4, 268–274. [Google Scholar] [CrossRef] [PubMed]
  130. Mays, K.S.; Lipman, J.J.; Schnapp, M. Local analgesia without anesthesia using peripheral perineural morphine injections. Anesth. Anal. 1987, 66, 417–420. [Google Scholar] [CrossRef] [PubMed]
  131. Cleary, J.; Mikus, G.; Somogyi, A.; Bochner, F. The influence of pharmacogenetics on opioid analgesia: Studies with codeine and oxycodone in the Sprague-Dawley/Dark Agouti rat model. J. Pharmacol. Exp. Ther. 1994, 271, 1528–1534. [Google Scholar] [PubMed]
  132. Mikus, G.; Somogyi, A.A.; Bochner, F.; Eichelbaum, M. Codeine O-demethylation: Rat strain differences and the effects of inhibitors. Biochem. Pharmacol. 1991, 41, 757–762. [Google Scholar] [CrossRef]
  133. Hermanns, H.; Hollmann, M.W.; Stevens, M.F.; Lirk, P.; Brandenburger, T.; Piegeler, T.; Werdehausen, R. Molecular mechanisms of action of systemic lidocaine in acute and chronic pain: A narrative review. Br. J. Anaesth. 2019, 123, 335–349. [Google Scholar] [CrossRef]
  134. Hille, B. Ionic Channels of Excitable Membranes; Sinauer Associates Inc.: Sunderland, MA, USA, 1984. [Google Scholar]
  135. Kirillova, I.; Teliban, A.; Gorodetskaya, N.; Grossmann, L.; Bartsch, F.; Rausch, V.H.; Struck, M.; Tode, J.; Baron, R.; Jänig, W. Effect of local and intravenous lidocaine on ongoing activity in injured afferent nerve fibers. Pain 2011, 152, 1562–1571. [Google Scholar] [CrossRef]
  136. Shin, J.W.; Pancaro, C.; Wang, C.F.; Gerner, P. Low-dose systemic bupivacaine prevents the development of allodynia after thoracotomy in rats. Anesth. Analg. 2008, 107, 1587–1591. [Google Scholar] [CrossRef]
  137. Delorme, C.; Navez, M.L.; Legout, V.; Deleens, R.; Moyse, D. Treatment of neuropathic pain with 5% lidocaine-medicated plaster: Five years of clinical experience. Pain Res. Manag. 2011, 16, 259–263. [Google Scholar] [CrossRef] [Green Version]
  138. Kalso, E.; Tramèr, M.R.; McQuay, H.J.; Moore, R.A. Systemic local-anaesthetic-type drugs in chronic pain: A systematic review. Eur. J. Pain 1998, 2, 3–14. [Google Scholar] [CrossRef]
  139. Tremont-Lukats, I.W.; Challapalli, V.; McNicol, E.D.; Lau, J.; Carr, D.B. Systemic administration of local anesthetics to relieve neuropathic pain: A systematic review and meta-analysis. Anesth. Analg. 2005, 101, 1738–1749. [Google Scholar] [CrossRef] [PubMed]
  140. Zhu, B.; Zhou, X.; Zhou, Q.; Wang, H.; Wang, S.; Luo, K. Intra-venous lidocaine to relieve neuropathic pain: A systematic review and meta-analysis. Front. Neurol. 2019, 10, 954. [Google Scholar] [CrossRef] [PubMed]
  141. Piao, L.-H.; Fujita, T.; Jiang, C.-Y.; Liu, T.; Yue, H.-Y.; Nakatsuka, T.; Kumamoto, E. TRPA1 activation by lidocaine in nerve terminals results in glutamate release increase. Biochem. Biophys. Res. Commun. 2009, 379, 980–984. [Google Scholar] [CrossRef]
  142. Piao, L.-H.; Fujita, T.; Yu, T.; Kumamoto, E. Presynaptic facilitation by tetracaine of glutamatergic spontaneous excitatory transmission in the rat spinal substantia gelatinosa—Involvement of TRPA1 channels. Brain Res. 2017, 1657, 245–252. [Google Scholar] [CrossRef] [PubMed]
  143. Leffler, A.; Fischer, M.J.; Rehner, D.; Kienel, S.; Kistner, K.; Sauer, S.K.; Gavva, N.R.; Reeh, P.W.; Nau, C. The vanilloid receptor TRPV1 is activated and sensitized by local anesthetics in rodent sensory neurons. J. Clin. Investig. 2008, 118, 763–776. [Google Scholar] [CrossRef] [PubMed]
  144. Leffler, A.; Lattrell, A.; Kronewald, S.; Niedermirtl, F.; Nau, C. Activation of TRPA1 by membrane permeable local anesthetics. Mol. Pain 2011, 7, 62. [Google Scholar] [CrossRef]
  145. Oda, A.; Iida, H.; Tanahashi, S.; Osawa, Y.; Yamaguchi, S.; Dohi, S. Effects of α2-adrenoceptor agonists on tetrodotoxin-resistant Na+ channels in rat dorsal root ganglion neurons. Eur. J. Anaesthesiol. 2007, 24, 934–941. [Google Scholar] [CrossRef]
  146. Zhao, K.; Dong, Y.; Su, G.; Wang, T.; Ji, T.; Wu, N.; Cui, X.; Li, W.; Yang, Y.; Chen, X. Effect of systemic lidocaine on postoperative early recovery quality in patients undergoing supratentorial tumor resection. Drug Des. Dev. Ther. 2022, 16, 1171–1181. [Google Scholar] [CrossRef]
  147. Chan, V.W.S.; Weisbrod, M.J.; Kaszas, Z.; Dragomir, C. Comparison of ropivacaine and lidocaine for intravenous regional anesthesia in volunteers: A preliminary study on anesthetic efficacy and blood level. Anesthesiology 1999, 90, 1602–1608. [Google Scholar] [CrossRef]
  148. McClellan, K.J.; Faulds, D. Ropivacaine: An update of its use in regional anaesthesia. Drugs 2000, 60, 1065–1093. [Google Scholar] [CrossRef]
  149. Bader, A.M.; Datta, S.; Flanagan, H.; Covino, B.G. Comparison of bupivacaine- and ropivacaine-induced conduction blockade in the isolated rabbit vagus nerve. Anesth. Analg. 1989, 68, 724–727. [Google Scholar] [CrossRef] [PubMed]
  150. Yilmaz-Rastoder, E.; Gold, M.S.; Hough, K.A.; Gebhart, G.F.; Williams, B.A. Effect of adjuvant drugs on the action of local anesthetics in isolated rat sciatic nerves. Reg. Anesth. Pain Med. 2012, 37, 403–409. [Google Scholar] [CrossRef] [PubMed]
  151. Lee-Son, S.; Wang, G.K.; Concus, A.; Crill, E.; Strichartz, G. Stereoselective inhibition of neuronal sodium channels by local anesthetics. Evidence for two sites of action? Anesthesiology 1992, 77, 324–335. [Google Scholar] [CrossRef]
  152. Foster, R.H.; Markham, A. Levobupivacaine: A review of its pharmacology and use as a local anaesthetic. Drugs 2000, 59, 551–579. [Google Scholar] [CrossRef] [PubMed]
  153. Vladimirov, M.; Nau, C.; Mok, W.M.; Strichartz, G. Potency of bupivacaine stereoisomers tested in vitro and in vivo: Biochemical, electrophysiological, and neurobehavioral studies. Anesthesiology 2000, 93, 744–755. [Google Scholar] [CrossRef] [Green Version]
  154. Stoetzer, C.; Martell, C.; de la Roche, J.; Leffler, A. Inhibition of voltage-gated Na+ channels by bupivacaine is enhanced by the adjuvants buprenorphine, ketamine, and clonidine. Reg. Anesth. Pain Med. 2017, 42, 462–468. [Google Scholar] [CrossRef] [PubMed]
  155. Gerner, P.; Mujtaba, M.; Sinnott, C.J.; Wang, G.K. Amitriptyline versus bupivacaine in rat sciatic nerve blockade. Anesthesiology 2001, 94, 661–667. [Google Scholar] [CrossRef] [Green Version]
  156. Bedford, J.A.; Turner, C.E.; Elsohly, H.N. Local anesthetic effects of cocaine and several extracts of the coca leaf (E. coca). Pharmacol. Biochem. Behav. 1984, 20, 819–821. [Google Scholar] [CrossRef]
  157. Pagala, M.K.D.; Venkatachari, S.A.T.; Herzlich, B.; Ravindran, K.; Namba, T.; Grob, D. Effect of cocaine on responses of mouse phrenic nerve-diaphragm preparation. Life Sci. 1991, 48, 795–802. [Google Scholar] [CrossRef]
  158. Carney, T.P. Alkaloids as local anesthetics. In The Alkaloids; Manske, R.H.F., Ed.; Academic Press: New York, NY, USA, 1955; Volume 5, pp. 211–227. [Google Scholar]
  159. Matthews, J.C.; Collins, A. Interactions of cocaine and cocaine congeners with sodium channels. Biochem. Pharmacol. 1983, 32, 455–460. [Google Scholar] [CrossRef]
  160. O’Leary, M.E.; Chahine, M. Cocaine binds to a common site on open and inactivated human heart (Nav1.5) sodium channels. J. Physiol. 2002, 541, 701–716. [Google Scholar] [CrossRef] [PubMed]
  161. Liu, D.; Hariman, R.J.; Bauman, J.L. Cocaine concentration-effect relationship in the presence and absence of lidocaine: Evidence of competitive binding between cocaine and lidocaine. J. Pharmacol. Exp. Ther. 1996, 276, 568–577. [Google Scholar] [PubMed]
  162. Chen, Y.-H.; Lin, C.-H.; Lin, P.-L.; Tsai, M.-C. Cocaine elicits action potential bursts in a central snail neuron: The role of delayed rectifying K+ current. Neuroscience 2006, 138, 257–280. [Google Scholar] [CrossRef] [PubMed]
  163. Štolc, S.; Mai, P.-M. Comparison of local anesthetic activity of pentacaine (trapencaine) and some of its derivatives by three different techniques. Pharmazie 1993, 48, 210–212. [Google Scholar] [PubMed]
  164. Tomohiro, D.; Mizuta, K.; Fujita, T.; Nishikubo, Y.; Kumamoto, E. Inhibition by capsaicin and its related vanilloids of compound action potentials in frog sciatic nerves. Life Sci. 2013, 92, 368–378. [Google Scholar] [CrossRef] [PubMed]
  165. Butterworth, J.F., IV; Lief, P.A.; Strichartz, G.R. The pH-dependent local anesthetic activity of diethylaminoethanol, a procaine metabolite. Anesthesiology 1988, 68, 501–506. [Google Scholar] [CrossRef]
  166. Ribeiro, J.A.; Sebastião, A.M. Antagonism of tetrodotoxin- and procaine-induced axonal blockade by adenine nucleotides in the frog sciatic nerve. Br. J. Pharmacol. 1984, 81, 277–282. [Google Scholar] [CrossRef] [Green Version]
  167. Kalichman, M.W.; Moorhouse, D.F.; Powell, H.C.; Myers, R.R. Relative neural toxicity of local anesthetics. J. Neuropathol. Exp. Neurol. 1993, 52, 234–240. [Google Scholar] [CrossRef]
  168. Lee, H.-S. Recent advances in topical anesthesia. J. Dent. Anesth. Pain Med. 2016, 16, 237–244. [Google Scholar] [CrossRef] [Green Version]
  169. Thygesen, M.M.; Rasmussen, M.M.; Madsen, J.G.; Pedersen, M.; Lauridsen, H. Propofol (2, 6-diisopropylphenol) is an applicable immersion anesthetic in the axolotl with potential uses in hemodynamic and neurophysiological experiments. Regeneration 2017, 4, 124–131. [Google Scholar] [CrossRef]
  170. Guénette, S.A.; Giroux, M.-C.; Vachon, P. Pain perception and anaesthesia in research frogs. Exp. Anim. 2013, 62, 87–92. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Vanable, J.W. Benzocaine: An excellent amphibian anesthetic. Axolotl Newsl. 1985, 14, 19–21. [Google Scholar]
  172. Starke, K.; Wagner, J.; Schümann, H.J. Adrenergic neuron blockade by clonidine: Comparison with guanethidine and local anesthetics. Arch. Int. Pharmacodyn. 1972, 195, 291–308. [Google Scholar] [PubMed]
  173. Gissen, A.J.; Covino, B.G.; Gregus, J. Differential sensitivities of mammalian nerve fibers to local anesthetic agents. Anesthesiology 1980, 53, 467–474. [Google Scholar] [CrossRef]
  174. Macdonald, R.L. Cellular effects of antiepileptic drugs. In Epilepsy: A Comprehensive Textbook; Engel, J., Jr., Pedley, T.A., Eds.; Lippincon-Raven Publishers: Philadelphia, PA, USA, 1997; pp. 1383–1391. [Google Scholar]
  175. Kubacka, M.; Rapacz, A.; Sałat, K.; Filipek, B.; Cios, A.; Pociecha, K.; Wyska, E.; Hubicka, U.; Żuromska-Witek, B.; Kwiecień, A.; et al. KM-416, a novel phenoxyalkylaminoalkanol derivative with anticonvulsant properties exerts analgesic, local anesthetic, and antidepressant-like activities. Pharmacodynamic, pharmacokinetic, and forced degradation studies. Eur. J. Pharmacol. 2020, 886, 173540. [Google Scholar] [CrossRef]
  176. Xie, X.; Dale, T.J.; John, V.H.; Cater, H.L.; Peakman, T.C.; Clare, J.J. Electrophysiological and pharmacological properties of the human brain type IIA Na+ channel expressed in a stable mammalian cell line. Pflüg. Arch. 2001, 441, 425–433. [Google Scholar] [CrossRef]
  177. McLean, M.J.; Macdonald, R.L. Carbamazepine and 10,11-epoxycarbamazepine produce use- and voltage-dependent limitation of rapidly firing action potentials of mouse central neurons in cell culture. J. Pharmacol. Exp. Ther. 1986, 238, 727–738. [Google Scholar]
  178. Cruccu, G.; Gronseth, G.; Alksne, J.; Argoff, C.; Brainin, M.; Burchiel, K.; Nurmikko, T.; Zakrzewska, J.M. AAN-EFNS guidelines on trigeminal neuralgia management. Eur. J. Neurol. 2008, 15, 1013–1028. [Google Scholar] [CrossRef]
  179. Vargas-Espinosa, M.-L.; Sanmartí-García, G.; Vázquez-Delgado, E.; Gay-Escoda, C. Antiepileptic drugs for the treatment of neuropathic pain: A systematic review. Med. Oral Patol. Oral Cir. Bucal 2012, 17, e786–e793. [Google Scholar] [CrossRef] [Green Version]
  180. Lang, D.G.; Wang, C.M.; Cooper, B.R. Lamotrigine, phenytoin and carbamazepine interactions on the sodium current present in N4TG1 mouse neuroblastoma cells. J. Pharmacol. Exp. Ther. 1993, 266, 829–835. [Google Scholar]
  181. Benes, J.; Parada, A.; Figueiredo, A.A.; Alves, P.C.; Freitas, A.P.; Learmonth, D.A.; Cunha, R.A.; Garrett, J.; Soares-da-Silva, P. Anticonvulsant and sodium channel-blocking properties of novel 10,11-dihydro-5H-dibenz[b,f]azepine-5-carboxamide derivatives. J. Med. Chem. 1999, 42, 2582–2587. [Google Scholar] [CrossRef] [PubMed]
  182. Huang, C.-W.; Huang, C.-C.; Lin, M.-W.; Tsai, J.-J.; Wu, S.-N. The synergistic inhibitory actions of oxcarbazepine on voltage-gated sodium and potassium currents in differentiated NG108-15 neuronal cells and model neurons. Int. J. Neuropsychopharmacol. 2008, 11, 597–610. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Kuo, C.-C. A common anticonvulsant binding site for phenytoin, carbamazepine, and lamotrigine in neuronal Na+ channels. Mol. Pharmacol. 1998, 54, 712–721. [Google Scholar]
  184. Molnár, P.; Erdö, S.L. Vinpocetine is as potent as phenytoin to block voltage-gated Na+ channels in rat cortical neurons. Eur. J. Pharmacol. 1995, 273, 303–306. [Google Scholar] [CrossRef] [PubMed]
  185. Qiao, X.; Sun, G.; Clare, J.J.; Werkman, T.R.; Wadman, W.J. Properties of human brain sodium channel α-subunits expressed in HEK293 cells and their modulation by carbamazepine, phenytoin and lamotrigine. Br. J. Pharmacol. 2014, 171, 1054–1067. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Neumcke, B.; Schwarz, J.R.; Stämpfli, R. A comparison of sodium currents in rat and frog myelinated nerve: Normal and modified sodium inactivation. J. Physiol. 1987, 382, 175–191. [Google Scholar] [CrossRef]
  187. Maneuf, Y.P.; Gonzalez, M.I.; Sutton, K.S.; Chung, F.-Z.; Pinnock, R.D.; Lee, K. Cellular and molecular action of the putative GABA-mimetic, gabapentin. Cell. Mol. Life Sci. 2003, 60, 742–750. [Google Scholar]
  188. Chen, J.; Li, L.; Chen, S.-R.; Chen, H.; Xie, J.-D.; Sirrieh, R.E.; MacLean, D.M.; Zhang, Y.; Zhou, M.-H.; Jayaraman, V.; et al. The α2δ-1-NMDA receptor complex is critically involved in neuropathic pain development and gabapentin therapeutic actions. Cell Rep. 2018, 22, 2307–2321. [Google Scholar] [CrossRef] [Green Version]
  189. Zona, C.; Ciotti, M.T.; Avoli, M. Topiramate attenuates voltage-gated sodium currents in rat cerebellar granule cells. Neurosci. Lett. 1997, 231, 123–126. [Google Scholar] [CrossRef]
  190. Curia, G.; Aracri, P.; Colombo, E.; Scalmani, P.; Mantegazza, M.; Avanzini, G.; Franceschetti, S. Phosphorylation of sodium channels mediated by protein kinase-C modulates inhibition by topiramate of tetrodotoxin-sensitive transient sodium current. Br. J. Pharmacol. 2007, 150, 792–797. [Google Scholar] [CrossRef] [Green Version]
  191. Chapman, A.; Keane, P.E.; Meldrum, B.S.; Simiand, J.; Vernieres, J.C. Mechanism of anticonvulsant action of valproate. Prog. Neurobiol. 1982, 19, 315–359. [Google Scholar] [PubMed]
  192. Perucca, E. A pharmacological and clinical review on topiramate, a new antiepileptic drug. Pharmacol. Res. 1997, 35, 241–256. [Google Scholar] [CrossRef] [PubMed]
  193. Braga, M.F.M.; Aroniadou-Anderjaska, V.; Li, H.; Rogawski, M.A. Topiramate reduces excitability in the basolateral amygdala by selectively inhibiting GluK1 (GluR5) kainate receptors on interneurons and positively modulating GABAA receptors on principal neurons. J. Pharmacol. Exp. Ther. 2009, 330, 558–566. [Google Scholar] [CrossRef] [PubMed]
  194. Lee, C.-Y.; Fu, W.-M.; Chen, C.-C.; Su, M.-J.; Liou, H.-H. Lamotrigine inhibits postsynaptic AMPA receptor and glutamate release in the dentate gyrus. Epilepsia 2008, 49, 888–897. [Google Scholar] [CrossRef] [PubMed]
  195. Blackburn-Munro, G.; Ibsen, N.; Erichsen, H.K. A comparison of the anti-nociceptive effects of voltage-activated Na+ channel blockers in the formalin test. Eur. J. Pharmacol. 2002, 445, 231–238. [Google Scholar] [CrossRef]
  196. Shannon, H.E.; Eberle, E.L.; Peters, S.C. Comparison of the effects of anticonvulsant drugs with diverse mechanisms of action in the formalin test in rats. Neuropharmacology 2005, 48, 1012–1020. [Google Scholar] [CrossRef]
  197. Douglas-Hall, P.; Dzahini, O.; Gaughran, F.; Bile, A.; Taylor, D. Variation in dose and plasma level of lamotrigine in patients discharged from a mental health trust. Ther. Adv. Psychopharmacol. 2017, 7, 17–24. [Google Scholar] [CrossRef] [Green Version]
  198. Morselli, P.L. Carbamazepine: Absorption, distribution, and excretion. In Antiepileptic Drugs, 4th ed.; Levy, R.H., Mattson, R.H., Meldrum, B.S., Eds.; Raven Press: New York, NY, USA, 1995; pp. 515–528. [Google Scholar]
  199. Ardid, D.; Jourdan, D.; Mestre, C.; Villanueva, L.; Le Bars, D.; Eschalier, A. Involvement of bulbospinal pathways in the antinociceptive effect of clomipramine in the rat. Brain Res. 1995, 695, 253–256. [Google Scholar] [CrossRef]
  200. Max, M.B.; Lynch, S.A.; Muir, J.; Shoaf, S.E.; Smoller, B.; Dubner, R. Effects of desipramine, amitriptyline, and fluoxetine on pain in diabetic neuropathy. N. Engl. J. Med. 1992, 326, 1250–1256. [Google Scholar] [CrossRef]
  201. Anjaneyulu, M.; Chopra, K. Possible involvement of cholinergic and opioid receptor mechanisms in fluoxetine mediated antinociception response in streptozotocin-induced diabetic mice. Eur. J. Pharmacol. 2006, 538, 80–84. [Google Scholar] [CrossRef]
  202. Cervantes-Durán, C.; Rocha-González, H.I.; Granados-Soto, V. Peripheral and spinal 5-HT receptors participate in the pronociceptive and antinociceptive effects of fluoxetine in rats. Neuroscience 2013, 252, 396–409. [Google Scholar] [CrossRef] [PubMed]
  203. Ghelardini, C.; Galeotti, N.; Bartolini, A. Antinociception induced by amitriptyline and imipramine is mediated by α2A-adrenoceptors. Jpn. J. Pharmacol. 2000, 82, 130–137. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Hall, H.; Ögren, S.-O. Effects of antidepressant drugs on different receptors in the brain. Eur. J. Pharmacol. 1981, 70, 393–407. [Google Scholar] [CrossRef] [PubMed]
  205. O’Donnell, J.M.; Shelton, R.C. Drug therapy of depression and anxiety disorders. In Goodman & Gilman’s the Pharmacological Basis of Therapeutics, 12th ed.; Brunton, L.L., Chabner, B.A., Knollmann, B.C., Eds.; McGraw-Hill, Medical Publishing Division: New York, NY, USA, 2011; pp. 397–415. [Google Scholar]
  206. Wong, D.T.; Bymaster, F.P. Dual serotonin and noradrenaline uptake inhibitor class of antidepressants—Potential for greater efficacy or just hype? Prog. Drug Res. 2002, 58, 169–222. [Google Scholar] [PubMed]
  207. Traboulsie, A.; Chemin, J.; Kupfer, E.; Nargeot, J.; Lory, P. T-type calcium channels are inhibited by fluoxetine and its metabolite norfluoxetine. Mol. Pharmacol. 2006, 69, 1963–1968. [Google Scholar] [CrossRef] [Green Version]
  208. Wu, W.; Ye, Q.; Wang, W.; Yan, L.; Wang, Q.; Xiao, H.; Wan, Q. Amitriptyline modulates calcium currents and intracellular calcium concentration in mouse trigeminal ganglion neurons. Neurosci. Lett. 2012, 506, 307–311. [Google Scholar] [CrossRef]
  209. Reynolds, I.J.; Miller, R.J. Tricyclic antidepressants block N-methyl-D-aspartate receptors: Similarities to the action of zinc. Br. J. Pharmacol. 1988, 95, 95–102. [Google Scholar] [CrossRef] [Green Version]
  210. Sernagor, E.; Kuhn, D.; Vyklicky, L., Jr.; Mayer, M.L. Open channel block of NMDA receptor responses evoked by tricyclic antidepressants. Neuron 1989, 2, 1221–1227. [Google Scholar] [CrossRef]
  211. Watanabe, Y.; Saito, H.; Abe, K. Tricyclic antidepressants block NMDA receptor-mediated synaptic responses and induction of long-term potentiation in rat hippocampal slices. Neuropharmacology 1993, 32, 479–486. [Google Scholar] [CrossRef]
  212. Barygin, O.I.; Nagaeva, E.I.; Tikhonov, D.B.; Belinskaya, D.A.; Vanchakova, N.P.; Shestakova, N.N. Inhibition of the NMDA and AMPA receptor channels by antidepressants and antipsychotics. Brain Res. 2017, 1660, 58–66. [Google Scholar] [CrossRef]
  213. Nagata, K.; Imai, T.; Yamashita, T.; Tsuda, M.; Tozaki-Saitoh, H.; Inoue, K. Antidepressants inhibit P2X4 receptor function: A possible involvement in neuropathic pain relief. Mol. Pain 2009, 5, 20. [Google Scholar] [CrossRef]
  214. Kremer, M.; Yalcin, I.; Goumon, Y.; Wurtz, X.; Nexon, L.; Daniel, D.; Megat, S.; Ceredig, R.A.; Ernst, C.; Turecki, G.; et al. A dual noradrenergic mechanism for the relief of neuropathic allodynia by the antidepressant drugs duloxetine and amitriptyline. J. Neurosci. 2018, 38, 9934–9954. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Le Cudennec, C.; Castagné, V. Face-to-face comparison of the predictive validity of two models of neuropathic pain in the rat: Analgesic activity of pregabalin, tramadol and duloxetine. Eur. J. Pharmacol. 2014, 735, 17–25. [Google Scholar] [CrossRef] [PubMed]
  216. Wong, D.T.; Bymaster, F.P.; Mayle, D.A.; Reid, L.R.; Krushinski, J.H.; Robertson, D.W. LY248686, a new inhibitor of serotonin and norepinephrine uptake. Neuropsychopharmacology 1993, 8, 23–33. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Russell, I.J.; Mease, P.J.; Smith, T.R.; Kajdasz, D.K.; Wohlreich, M.M.; Detke, M.J.; Walker, D.J.; Chappell, A.S.; Arnold, L.M. Efficacy and safety of duloxetine for treatment of fibromyalgia in patients with or without major depressive disorder: Results from a 6-month, randomized, double-blind, placebo-controlled, fixed-dose trial. Pain 2008, 136, 432–444. [Google Scholar] [CrossRef]
  218. Müller, N.; Schennach, R.; Riedel, M.; Möller, H.-J. Duloxetine in the treatment of major psychiatric and neuropathic disorders. Expert Rev. Neurother. 2008, 8, 527–536. [Google Scholar] [CrossRef]
  219. Stark, P.; Fuller, R.W.; Wong, D.T. The pharmacologic profile of fluoxetine. J. Clin. Psychiatry 1985, 46, 7–13. [Google Scholar]
  220. Korzeniewska-Rybicka, I.; Płaźnik, A. Analgesic effect of antidepressant drugs. Pharmacol. Biochem. Behav. 1998, 59, 331–338. [Google Scholar] [CrossRef]
  221. Richeimer, S.H.; Bajwa, Z.H.; Kahraman, S.S.; Ransil, B.J.; Warfield, C.A. Utilization patterns of tricyclic antidepressants in a multidisciplinary pain clinic: A survey. Clin. J. Pain 1997, 13, 324–329. [Google Scholar] [CrossRef]
  222. Okuda, K.; Takanishi, T.; Yoshimoto, K.; Ueda, S. Trazodone hydrochloride attenuates thermal hyperalgesia in a chronic constriction injury rat model. Eur. J. Anaesthesiol. 2003, 20, 409–415. [Google Scholar] [CrossRef]
  223. Richelson, E.; Pfenning, M. Blockade by antidepressants and related compounds of biogenic amine uptake into rat brain synaptosomes: Most antidepressants selectively block norepinephrine uptake. Eur. J. Pharmacol. 1984, 104, 277–286. [Google Scholar] [CrossRef] [PubMed]
  224. Schreiber, S.; Backer, M.M.; Herman, I.; Shamir, D.; Boniel, T.; Pick, C.G. The antinociceptive effect of trazodone in mice is mediated through both µ-opioid and serotonergic mechanisms. Behav. Brain Res. 2000, 114, 51–56. [Google Scholar] [CrossRef] [PubMed]
  225. Davidoff, G.; Guarracini, M.; Roth, E.; Sliwa, J.; Yarkony, G. Trazodone hydrochloride in the treatment of dysesthetic pain in traumatic myelopathy: A randomized, double-blind, placebo-controlled study. Pain 1987, 29, 151–161. [Google Scholar] [CrossRef] [PubMed]
  226. Baastrup, C.; Finnerup, N.B. Pharmacological management of neuropathic pain following spinal cord injury. CNS Drugs 2008, 22, 455–475. [Google Scholar] [CrossRef]
  227. Stoetzer, C.; Papenberg, B.; Doll, T.; Völker, M.; Heineke, J.; Stoetzer, M.; Wegner, F.; Leffler, A. Differential inhibition of cardiac and neuronal Na+ channels by the selective serotonin-norepinephrine reuptake inhibitors duloxetine and venlafaxine. Eur. J. Pharmacol. 2016, 783, 1–10. [Google Scholar] [CrossRef]
  228. Wang, S.-Y.; Calderon, J.; Wang, G.K. Block of neuronal Na+ channels by antidepressant duloxetine in a state-dependent manner. Anesthesiology 2010, 113, 655–665. [Google Scholar] [CrossRef] [Green Version]
  229. Pancrazio, J.J.; Kamatchi, G.L.; Roscoe, A.K.; Lynch, C., 3rd. Inhibition of neuronal Na+ channels by antidepressant drugs. J. Pharmacol. Exp. Ther. 1998, 284, 208–214. [Google Scholar]
  230. Ishii, Y.; Sumi, T. Amitriptyline inhibits striatal efflux of neurotransmitters via blockade of voltage-dependent Na+ channels. Eur. J. Pharmacol. 1992, 221, 377–380. [Google Scholar] [CrossRef]
  231. Leffler, A.; Reiprich, A.; Mohapatra, D.P.; Nau, C. Use-dependent block by lidocaine but not amitriptyline is more pronounced in tetrodotoxin (TTX)-resistant Nav1.8 than in TTX-sensitive Na+ channels. J. Pharmacol. Exp. Ther. 2007, 320, 354–364. [Google Scholar] [CrossRef] [Green Version]
  232. Nicholson, G.M.; Blanche, T.; Mansfield, K.; Tran, Y. Differential blockade of neuronal voltage-gated Na+ and K+ channels by antidepressant drugs. Eur. J. Pharmacol. 2002, 452, 35–48. [Google Scholar] [CrossRef]
  233. Song, J.-H.; Ham, S.-S.; Shin, Y.-K.; Lee, C.-S. Amitriptyline modulation of Na+ channels in rat dorsal root ganglion neurons. Eur. J. Pharmacol. 2000, 401, 297–305. [Google Scholar] [CrossRef] [PubMed]
  234. Wang, G.K.; Russell, C.; Wang, S.-Y. State-dependent block of voltage-gated Na+ channels by amitriptyline via the local anesthetic receptor and its implication for neuropathic pain. Pain 2004, 110, 166–174. [Google Scholar] [CrossRef] [PubMed]
  235. Yan, L.; Wang, Q.; Fu, Q.; Ye, Q.; Xiao, H.; Wan, Q. Amitriptyline inhibits currents and decreases the mRNA expression of voltage-gated sodium channels in cultured rat cortical neurons. Brain Res. 2010, 1336, 1–9. [Google Scholar] [CrossRef] [PubMed]
  236. Dick, I.E.; Brochu, R.M.; Purohit, Y.; Kaczorowski, G.J.; Martin, W.J.; Priest, B.T. Sodium channel blockade may contribute to the analgesic efficacy of antidepressants. J. Pain 2007, 8, 315–324. [Google Scholar] [CrossRef] [PubMed]
  237. Liang, J.; Liu, X.; Pan, M.; Dai, W.; Dong, Z.; Wang, X.; Liu, R.; Zheng, J.; Yu, S. Blockade of Nav1.8 currents in nociceptive trigeminal neurons contributes to anti-trigeminovascular nociceptive effect of amitriptyline. Neuromol. Med. 2014, 16, 308–321. [Google Scholar] [CrossRef]
  238. Catterall, W.A.; Mackie, K. Local anesthetics. In Goodman & Gilman’s the Pharmacological Basis of Therapeutics, 12th ed.; Brunton, L.L., Chabner, B.A., Knollmann, B.C., Eds.; McGraw-Hill, Medical Publishing Division: New York, NY, USA, 2011; pp. 565–582. [Google Scholar]
  239. Caruso, R.; Ostuzzi, G.; Turrini, G.; Ballette, F.; Recla, E.; Dall’Olio, R.; Croce, E.; Casoni, B.; Grassi, L.; Barbui, C. Beyond pain: Can antidepressants improve depressive symptoms and quality of life in patients with neuropathic pain? A systematic review and meta-analysis. Pain 2019, 160, 2186–2198. [Google Scholar] [CrossRef]
  240. Mika, J.; Zychowska, M.; Makuch, W.; Rojewska, E.; Przewlocka, B. Neuronal and immunological basis of action of antidepressants in chronic pain—Clinical and experimental studies. Pharmacol. Rep. 2013, 65, 1611–1621. [Google Scholar] [CrossRef]
  241. Banafshe, H.R.; Hajhashemi, V.; Minaiyan, M.; Mesdaghinia, A.; Abed, A. Antinociceptive effects of maprotiline in a rat model of peripheral neuropathic pain: Possible involvement of opioid system. Iran. J. Basic Med. Sci. 2015, 18, 752–757. [Google Scholar] [CrossRef]
  242. Bhana, N.; Goa, K.L.; McClellan, K.J. Dexmedetomidine. Drugs 2000, 59, 263–268. [Google Scholar] [CrossRef]
  243. Costa-Pereira, J.T.; Ribeiro, J.; Martins, I.; Tavares, I. Role of spinal cord α2-adrenoreceptors in noradrenergic inhibition of nociceptive transmission during chemotherapy-induced peripheral neuropathy. Front. Neurosci. 2020, 13, 1413. [Google Scholar] [CrossRef] [Green Version]
  244. Fisher, B.; Zornow, M.H.; Yaksh, T.L.; Peterson, B.M. Antinociceptive properties of intrathecal dexmedetomidine in rats. Eur. J. Pharmacol. 1991, 192, 221–225. [Google Scholar] [CrossRef] [PubMed]
  245. Takano, Y.; Yaksh, T.L. Relative efficacy of spinal alpha-2 agonists, dexmedetomidine, clonidine and ST-91, determined in vivo by using N-ethoxycarbonyl-2-ethoxy-1,2-dihydroquinoline, an irreversible antagonist. J. Pharmacol. Exp. Ther. 1991, 258, 438–446. [Google Scholar] [PubMed]
  246. Filos, K.S.; Goudas, L.C.; Patroni, O.; Polyzou, V. Hemodynamic and analgesic profile after intrathecal clonidine in humans. A dose-response study. Anesthesiology 1994, 81, 591–601. [Google Scholar] [CrossRef] [PubMed]
  247. Sullivan, A.F.; Kalso, E.A.; McQuay, H.J.; Dickenson, A.H. The antinociceptive actions of dexmedetomidine on dorsal horn neuronal responses in the anaesthetized rat. Eur. J. Pharmacol. 1992, 215, 127–133. [Google Scholar] [CrossRef] [PubMed]
  248. Brummett, C.M.; Norat, M.A.; Palmisano, J.M.; Lydic, R. Perineural administration of dexmedetomidine in combination with bupivacaine enhances sensory and motor blockade in sciatic nerve block without inducing neurotoxicity in rat. Anesthesiology 2008, 109, 502–511. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  249. Calasans-Maia, J.A.; Zapata-Sudo, G.; Sudo, R.T. Dexmedetomidine prolongs spinal anaesthesia induced by levobupivacaine 0.5% in guinea-pigs. J. Pharm. Pharmacol. 2005, 57, 1415–1420. [Google Scholar] [CrossRef]
  250. Tsutsui, Y.; Sunada, K. Adding dexmedetomidine to articaine increases the latency of thermal antinociception in rats. Anesth. Prog. 2020, 67, 72–78. [Google Scholar] [CrossRef]
  251. Marolf, V.; Ida, K.K.; Siluk, D.; Struck-Lewicka, W.; Markuszewski, M.J.; Sandersen, C. Effects of perineural administration of ropivacaine combined with perineural or intravenous administration of dexmedetomidine for sciatic and saphenous nerve blocks in dogs. Am. J. Vet. Res. 2021, 82, 449–458. [Google Scholar] [CrossRef]
  252. Kanazi, G.E.; Aouad, M.T.; Jabbour-Khoury, S.I.; Al Jazzar, M.D.; Alameddine, M.M.; Al-Yaman, R.; Bulbul, M.; Baraka, A.S. Effect of low-dose dexmedetomidine or clonidine on the characteristics of bupivacaine spinal block. Acta Anaesthesiol. Scand. 2006, 50, 222–227. [Google Scholar] [CrossRef]
  253. Madan, R.; Bharti, N.; Shende, D.; Khokhar, S.K.; Kaul, H.L. A dose response study of clonidine with local anesthetic mixture for peribulbar block: A comparison of three doses. Anesth. Analg. 2001, 93, 1593–1597. [Google Scholar] [CrossRef]
  254. Memiş, D.; Turan, A.; Karamanlioĝlu, B.; Pamukçu, Z.; Kurt, I. Adding dexmedetomidine to lidocaine for intravenous regional anesthesia. Anesth. Analg. 2004, 98, 835–840. [Google Scholar] [CrossRef] [PubMed]
  255. Singelyn, F.J.; Gouverneur, J.-M.; Robert, A. A minimum dose of clonidine added to mepivacaine prolongs the duration of anesthesia and analgesia after axillary brachial plexus block. Anesth. Analg. 1996, 83, 1046–1050. [Google Scholar] [CrossRef] [PubMed]
  256. Tschernko, E.M.; Klepetko, H.; Gruber, E.; Kritzinger, M.; Klimscha, W.; Jandrasits, O.; Haider, W. Clonidine added to the anesthetic solution enhances analgesia and improves oxygenation after intercostal nerve block for thoracotomy. Anesth. Analg. 1998, 87, 107–111. [Google Scholar] [CrossRef]
  257. Mohyiedin, H.; Kamelia, A.A.; Ekram, F.S.; Al Shaimaa, A.K. The effect of various additives to local anesthetics on the duration of analgesia of supraclavicular brachial plexus block. J. Anesth. Intensive Care Med. 2019, 9, 555756. [Google Scholar]
  258. Ouchi, K. Dexmedetomidine 2 ppm is appropriate for the enhancement effect of local anesthetic action of lidocaine in inferior alveolar nerve block: A preliminary, randomized cross-over study. Clin. J. Pain 2020, 36, 618–625. [Google Scholar] [CrossRef]
  259. Sane, S.; Shokouhi, S.; Golabi, P.; Rezaeian, M.; Kazemi Haki, B. The effect of dexmedetomidine in combination with bupivacaine on sensory and motor block time and pain score in supraclavicular block. Pain Res. Manag. 2021, 2021, 8858312. [Google Scholar] [CrossRef]
  260. Eisenach, J.C.; de Kock, M.; Klimscha, W. α2-Adrenergic agonists for regional anesthesia. A clinical review of clonidine (1984–1995). Anesthesiology 1996, 85, 655–674. [Google Scholar] [CrossRef]
  261. Concepcion, M.; Maddi, R.; Francis, D.; Rocco, A.G.; Murray, E.; Covino, B.G. Vasoconstrictors in spinal anesthesia with tetracaine—A comparison of epinephrine and phenylephrine. Anesth. Analg. 1984, 63, 134–138. [Google Scholar] [CrossRef]
  262. Vaida, G.T.; Moss, P.; Capan, L.M.; Turndorf, H. Prolongation of lidocaine spinal anesthesia with phenylephrine. Anesth. Analg. 1986, 65, 781–785. [Google Scholar] [CrossRef]
  263. Gaumann, D.M.; Brunet, P.C.; Jirounek, P. Clonidine enhances the effects of lidocaine on C-fiber action potential. Anesth. Analg. 1992, 74, 719–725. [Google Scholar] [CrossRef]
  264. Kawasaki, Y.; Kumamoto, E.; Furue, H.; Yoshimura, M. α2 Adrenoceptor-mediated presynaptic inhibition of primary afferent glutamatergic transmission in rat substantia gelatinosa neurons. Anesthesiology 2003, 98, 682–689. [Google Scholar] [CrossRef] [PubMed]
  265. Pan, Y.-Z.; Li, D.-P.; Pan, H.-L. Inhibition of glutamatergic synaptic input to spinal lamina IIo neurons by presynaptic α2-adrenergic receptors. J. Neurophysiol. 2002, 87, 1938–1947. [Google Scholar] [CrossRef] [PubMed]
  266. Butterworth, J.F., IV; Strichartz, G.R. The α2-adrenergic agonists clonidine and guanfacine produce tonic and phasic block of conduction in rat sciatic nerve fibers. Anesth. Analg. 1993, 76, 295–301. [Google Scholar] [PubMed]
  267. Ishii, H.; Kohno, T.; Yamakura, T.; Ikoma, M.; Baba, H. Action of dexmedetomidine on the substantia gelatinosa neurons of the rat spinal cord. Eur. J. Neurosci. 2008, 27, 3182–3190. [Google Scholar] [CrossRef] [PubMed]
  268. Yoshitomi, T.; Kohjitani, A.; Maeda, S.; Higuchi, H.; Shimada, M.; Miyawaki, T. Dexmedetomidine enhances the local anesthetic action of lidocaine via an α-2A adrenoceptor. Anesth. Analg. 2008, 107, 96–101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  269. Mohamed, S.A.; Sayed, D.M.; El Sherif, F.A.; Abd El-Rahman, A.M. Effect of local wound infiltration with ketamine versus dexmedetomidine on postoperative pain and stress after abdominal hysterectomy, a randomized trial. Eur. J. Pain 2018, 22, 951–960. [Google Scholar] [CrossRef]
  270. Coughlan, M.G.; Lee, J.G.; Bosnjak, Z.J.; Schmeling, W.T.; Kampine, J.P.; Warltier, D.C. Direct coronary and cerebral vascular responses to dexmedetomidine. Significance of endogenous nitric oxide synthesis. Anesthesiology 1992, 77, 998–1006. [Google Scholar] [CrossRef]
  271. Correa-Sales, C.; Rabin, B.C.; Maze, M. A hypnotic response to dexmedetomidine, an α2 agonist, is mediated in the locus coeruleus in rats. Anesthesiology 1992, 76, 948–952. [Google Scholar] [CrossRef]
  272. Sjöholm, B.; Voutilainen, R.; Luomala, K.; Savola, J.-M.; Scheinin, M. Characterization of [3H]atipamezole as a radioligand for α2-adrenoceptors. Eur. J. Pharmacol. 1992, 215, 109–117. [Google Scholar] [CrossRef]
  273. MacDonald, E.; Kobilka, B.K.; Scheinin, M. Gene targeting-homing in on α2-adrenoceptor-subtype function. Trends Pharmacol. Sci. 1997, 18, 211–219. [Google Scholar] [CrossRef]
  274. Virtanen, R. Pharmacological profiles of medetomidine and its antagonist, atipamezole. Acta Vet. Scand. 1989, 85, 29–37. [Google Scholar]
  275. Bylund, D.B.; Eikenberg, D.C.; Hieble, J.P.; Langer, S.Z.; Lefkowitz, R.J.; Minneman, K.P.; Molinoff, P.B.; Ruffolo, R.R.; Trendelenburg, U. International union of pharmacology nomenclature of adrenoceptors. Pharmacol. Rev. 1994, 46, 121–136. [Google Scholar] [PubMed]
  276. Chen, B.-S.; Peng, H.; Wu, S.-N. Dexmedetomidine, an α2-adrenergic agonist, inhibits neuronal delayed-rectifier potassium current and sodium current. Br. J. Anaesth. 2009, 103, 244–254. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  277. Ebert, T.J.; Hall, J.E.; Barney, J.A.; Uhrich, T.D.; Colinco, M.D. The effects of increasing plasma concentrations of dexmedetomidine in humans. Anesthesiology 2000, 93, 382–394. [Google Scholar] [CrossRef]
  278. Slingsby, L.S.; Taylor, P.M. Thermal antinociception after dexmedetomidine administration in cats: A dose-finding study. J. Vet. Pharmacol. Ther. 2008, 31, 135–142. [Google Scholar] [CrossRef]
  279. Bharti, N.; Sardana, D.K.; Bala, I. The analgesic efficacy of dexmedetomidine as an adjunct to local anesthetics in supraclavicular brachial plexus block: A randomized controlled trial. Anesth. Analg. 2015, 121, 1655–1660. [Google Scholar] [CrossRef]
  280. Nassar, M.A.; Stirling, L.C.; Forlani, G.; Baker, M.D.; Matthews, E.A.; Dickenson, A.H.; Wood, J.N. Nociceptor-specific gene deletion reveals a major role for Nav1.7 (PN1) in acute and inflammatory pain. Proc. Natl. Acad. Sci. USA 2004, 101, 12706–12711. [Google Scholar] [CrossRef] [Green Version]
  281. Patapoutian, A.; Tate, S.; Woolf, C.J. Transient receptor potential channels: Targeting pain at the source. Nat. Rev. Drug Discov. 2009, 8, 55–68. [Google Scholar] [CrossRef] [Green Version]
  282. Julius, D. TRP channels and pain. Annu. Rev. Cell Dev. Biol. 2013, 29, 355–384. [Google Scholar] [CrossRef] [Green Version]
  283. Wang, C.; Fujita, T.; Yasuda, H.; Kumamoto, E. Spontaneous excitatory transmission enhancement produced by linalool and its isomer geraniol in rat spinal substantia gelatinosa neurons—Involvement of transient receptor potential channels. Phytomed. Plus 2022, 2, 100155. [Google Scholar] [CrossRef]
  284. Kumamoto, E.; Fujita, T.; Jiang, C.-Y. TRP channels involved in spontaneous L-glutamate release enhancement in the adult rat spinal substantia gelatinosa. Cells 2014, 3, 331–362. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  285. Kumamoto, E.; Fujita, T. Differential activation of TRP channels in the adult rat spinal substantia gelatinosa by stereoisomers of plant-derived chemicals. Pharmaceuticals 2016, 9, 46. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  286. Guimarães, A.G.; Quintans, J.S.S.; Quintans-Júnior, L.J. Monoterpenes with analgesic activity—A systematic review. Phytother. Res. 2013, 27, 1–15. [Google Scholar] [CrossRef] [PubMed]
  287. Tsuchiya, H. Anesthetic agents of plant origin: A review of phytochemicals with anesthetic activity. Molecules 2017, 22, 1369. [Google Scholar] [CrossRef] [Green Version]
  288. Nolano, M.; Simone, D.A.; Wendelschafer-Crabb, G.; Johnson, T.; Hazen, E.; Kennedy, W.R. Topical capsaicin in humans: Parallel loss of epidermal nerve fibers and pain sensation. Pain 1999, 81, 135–145. [Google Scholar] [CrossRef]
  289. Malmberg, A.B.; Mizisin, A.P.; Calcutt, N.A.; von Stein, T.; Robbins, W.R.; Bley, K.R. Reduced heat sensitivity and epidermal nerve fiber immunostaining following single applications of a high-concentration capsaicin patch. Pain 2004, 111, 360–367. [Google Scholar] [CrossRef] [PubMed]
  290. Kawasaki, H.; Mizuta, K.; Fujita, T.; Kumamoto, E. Inhibition by menthol and its related chemicals of compound action potentials in frog sciatic nerves. Life Sci. 2013, 92, 359–367. [Google Scholar] [CrossRef]
  291. Matsushita, A.; Ohtsubo, S.; Fujita, T.; Kumamoto, E. Inhibition by TRPA1 agonists of compound action potentials in the frog sciatic nerve. Biochem. Biophys. Res. Commun. 2013, 434, 179–184. [Google Scholar] [CrossRef]
  292. Ohtsubo, S.; Fujita, T.; Matsushita, A.; Kumamoto, E. Inhibition of the compound action potentials of frog sciatic nerves by aroma oil compounds having various chemical structures. Pharmacol. Res. Perspect. 2015, 3, e00127. [Google Scholar] [CrossRef]
  293. Lundbæk, J.A.; Birn, P.; Tape, S.E.; Toombes, G.E.S.; Søgaard, R.; Koeppe II, R.E.; Gruner, S.M.; Hansen, A.J.; Andersen, O.S. Capsaicin regulates voltage-dependent sodium channels by altering lipid bilayer elasticity. Mol. Pharmacol. 2005, 68, 680–689. [Google Scholar] [CrossRef] [Green Version]
  294. Cao, X.; Cao, X.; Xie, H.; Yang, R.; Lei, G.; Li, F.; Li, A.; Liu, C.; Liu, L. Effects of capsaicin on VGSCs in TRPV1−/− mice. Brain Res. 2007, 1163, 33–43. [Google Scholar] [CrossRef] [PubMed]
  295. Wang, S.-Y.; Mitchell, J.; Wang, G.K. Preferential block of inactivation-deficient Na+ currents by capsaicin reveals a non-TRPV1 receptor within the Na+ channel. Pain 2007, 127, 73–83. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  296. Cho, J.S.; Kim, T.H.; Lim, J.-M.; Song, J.-H. Effects of eugenol on Na+ currents in rat dorsal root ganglion neurons. Brain Res. 2008, 1243, 53–62. [Google Scholar] [CrossRef] [PubMed]
  297. Haeseler, G.; Maue, D.; Grosskreutz, J.; Bufler, J.; Nentwig, B.; Piepenbrock, S.; Dengler, R.; Leuwer, M. Voltage-dependent block of neuronal and skeletal muscle sodium channels by thymol and menthol. Eur. J. Anaesthesiol. 2002, 19, 571–579. [Google Scholar] [CrossRef] [PubMed]
  298. Joca, H.C.; Cruz-Mendes, Y.; Oliveira-Abreu, K.; Maia-Joca, R.P.M.; Barbosa, R.; Lemos, T.L.; Lacerda Beirão, P.S.; Leal-Cardoso, J.H. Carvacrol decreases neuronal excitability by inhibition of voltage-gated sodium channels. J. Nat. Prod. 2012, 75, 1511–1517. [Google Scholar] [CrossRef]
  299. Leal-Cardoso, J.H.; da Silva-Alves, K.S.; Ferreira-da-Silva, F.W.; dos Santos-Nascimento, T.; Joca, H.C.; de Macedo, F.H.P.; de Albuquerque-Neto, P.M.; Magalhães, P.J.C.; Lahlou, S.; Cruz, J.S.; et al. Linalool blocks excitability in peripheral nerves and voltage-dependent Na+ current in dissociated dorsal root ganglia neurons. Eur. J. Pharmacol. 2010, 645, 86–93. [Google Scholar] [CrossRef]
  300. De Araújo, D.A.M.; Freitas, C.; Cruz, J.S. Essential oils components as a new path to understand ion channel molecular pharmacology. Life Sci. 2011, 89, 540–544. [Google Scholar] [CrossRef]
  301. Joca, H.C.; Vieira, D.C.O.; Vasconcelos, A.P.; Araújo, D.A.M.; Cruz, J.S. Carvacrol modulates voltage-gated sodium channels kinetics in dorsal root ganglia. Eur. J. Pharmacol. 2015, 756, 22–29. [Google Scholar] [CrossRef]
  302. Nozoe, T. Über die farbstoffe im holzteile des “hinoki” baumes. I. Hinokitin und hinokitiol. Bull. Chem. Soc. Jpn. 1936, 11, 295–298. [Google Scholar] [CrossRef]
  303. Magori, N.; Fujita, T.; Kumamoto, E. Hinokitiol inhibits compound action potentials in the frog sciatic nerve. Eur. J. Pharmacol. 2018, 819, 254–260. [Google Scholar] [CrossRef]
  304. Baba, T.; Nakano, H.; Tamai, K.; Sawamura, D.; Hanada, K.; Hashimoto, I.; Arima, Y. Inhibitory effect of β-thujaplicin on ultraviolet B-induced apoptosis in mouse keratinocytes. J. Investig. Dermatol. 1998, 110, 24–28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  305. Shih, Y.-H.; Lin, D.-J.; Chang, K.-W.; Hsia, S.-M.; Ko, S.-Y.; Lee, S.-Y.; Hsue, S.-S.; Wang, T.-H.; Chen, Y.-L.; Shieh, T.-M. Evaluation physical characteristics and comparison antimicrobial and anti-inflammation potentials of dental root canal sealers containing hinokitiol in vitro. PLoS ONE 2014, 9, e94941. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  306. Morita, Y.; Matsumura, E.; Okabe, T.; Shibata, M.; Sugiura, M.; Ohe, T.; Tsujibo, H.; Ishida, N.; Inamori, Y. Biological activity of tropolone. Biol. Pharm. Bull. 2003, 26, 1487–1490. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  307. Morita, Y.; Matsumura, E.; Okabe, T.; Fukui, T.; Ohe, T.; Ishida, N.; Inamori, Y. Biological activity of β-dolabrin, γ-thujaplicin, and 4-acetyltropolone, hinokitiol-related compounds. Biol. Pharm. Bull. 2004, 27, 1666–1669. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  308. Yamato, M.; Ando, J.; Sakaki, K.; Hashigaki, K.; Wataya, Y.; Tsukagoshi, S.; Tashiro, T.; Tsuruo, T. Synthesis and antitumor activity of tropolone derivatives. 7. Bistropolones containing connecting methylene chains. J. Med. Chem. 1992, 35, 267–273. [Google Scholar] [CrossRef] [PubMed]
  309. Inamori, Y.; Tsujibo, H.; Ohishi, H.; Ishii, F.; Mizugaki, M.; Aso, H.; Ishida, N. Cytotoxic effect of hinokitiol and tropolone on the growth of mammalian cells and on blastogenesis of mouse splenic T cells. Biol. Pharm. Bull. 1993, 16, 521–523. [Google Scholar] [CrossRef] [Green Version]
  310. Yasumoto, E.; Nakano, K.; Nakayachi, T.; Morshed, S.R.; Hashimoto, K.; Kikuchi, H.; Nishikawa, H.; Kawase, M.; Sakagami, H. Cytotoxic activity of deferiprone, maltol and related hydroxyketones against human tumor cell lines. Anticancer Res. 2004, 24, 755–762. [Google Scholar]
  311. Nagao, Y.; Sata, M. Effect of oral care gel on the quality of life for oral lichen planus in patients with chronic HCV infection. Virol. J. 2011, 8, 348. [Google Scholar] [CrossRef] [Green Version]
  312. James, R.; Glen, J.B. Synthesis, biological evaluation, and preliminary structure-activity considerations of a series of alkylphenols as intravenous anesthetic agents. J. Med. Chem. 1980, 23, 1350–1357. [Google Scholar] [CrossRef]
  313. Doze, V.A.; Westphal, L.M.; White, P.F. Comparison of propofol with methohexital for outpatient anesthesia. Anesth. Analg. 1986, 65, 1189–1195. [Google Scholar] [CrossRef]
  314. Shafer, A.; Doze, V.A.; Shafer, S.L.; White, P.F. Pharmacokinetics and pharmacodynamics of propofol infusions during general anesthesia. Anesthesiology 1988, 69, 348–356. [Google Scholar] [CrossRef] [PubMed]
  315. Antkowiak, B.; Rammes, G. GABA(A) receptor-targeted drug development—New perspectives in perioperative anesthesia. Expert Opin. Drug Discov. 2019, 14, 683–699. [Google Scholar] [CrossRef] [PubMed]
  316. Vasileiou, I.; Xanthos, T.; Koudouna, E.; Perrea, D.; Klonaris, C.; Katsargyris, A.; Papadimitriou, L. Propofol: A review of its non-anaesthetic effects. Eur. J. Pharmacol. 2009, 605, 1–8. [Google Scholar] [CrossRef] [PubMed]
  317. Hanrahan, S.J.; Greger, B.; Parker, R.A.; Ogura, T.; Obara, S.; Egan, T.D.; House, P.A. The effects of propofol on local field potential spectra, action potential firing rate, and their temporal relationship in humans and felines. Front. Hum. Neurosci. 2013, 7, 136. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  318. Shi, Q.-Q.; Sun, X.; Fang, H. A mechanism study on propofol’s action on middle latency auditory evoked potential by neurons in ventral partition of medial geniculate body in rats. Eur. Rev. Med. Pharmacol. Sci. 2014, 18, 1859–1868. [Google Scholar] [PubMed]
  319. Takazawa, T.; Furue, H.; Nishikawa, K.; Uta, D.; Takeshima, K.; Goto, F.; Yoshimura, M. Actions of propofol on substantia gelatinosa neurones in rat spinal cord revealed by in vitro and in vivo patch-clamp recordings. Eur. J. Neurosci. 2009, 29, 518–528. [Google Scholar] [CrossRef] [PubMed]
  320. Hijikata, Y. Analgesic treatment with Kampo prescription. Expert Rev. Neurother. 2006, 6, 795–802. [Google Scholar] [CrossRef]
  321. Kono, T.; Kanematsu, T.; Kitajima, M. Exodus of Kampo, traditional Japanese medicine, from the complementary and alternative medicines: Is it time yet? Surgery 2009, 146, 837–840. [Google Scholar] [CrossRef] [PubMed]
  322. Motoo, Y.; Arai, I.; Hyodo, I.; Tsutani, K. Current status of Kampo (Japanese herbal) medicines in Japanese clinical practice guidelines. Complement. Ther. Med. 2009, 17, 147–154. [Google Scholar] [CrossRef]
  323. Mochiki, E.; Yanai, M.; Ohno, T.; Kuwano, H. The effect of traditional Japanese medicine (Kampo) on gastrointestinal function. Surg. Today 2010, 40, 1105–1111. [Google Scholar] [CrossRef]
  324. Wachtel-Galor, S.; Benzie, I.F.F. Herbal medicine: An introduction to its history, usage, regulation, current trends, and research needs. In Herbal Medicine: Biomolecular and Clinical Aspects, 2nd ed.; Chapter 1; Benzie, I.F.F., Wachtel-Galor, S., Eds.; CRC Press: Boca Raton, FL, USA, 2011. [Google Scholar]
  325. Sunagawa, M.; Takayama, Y.; Kato, M.; Tanaka, M.; Fukuoka, S.; Okumo, T.; Tsukada, M.; Yamaguchi, K. Kampo formulae for the treatment of neuropathic pain—Especially the mechanism of action of Yokukansan. Front. Mol. Neurosci. 2021, 14, 705023. [Google Scholar] [CrossRef]
  326. Matsushita, A.; Fujita, T.; Ohtsubo, S.; Kumamoto, E. Traditional Japanese medicines inhibit compound action potentials in the frog sciatic nerve. J. Ethnopharmacol. 2016, 178, 272–280. [Google Scholar] [CrossRef]
  327. Fink, B.R.; Cairns, A.M. Differential slowing and block of conduction by lidocaine in individual afferent myelinated and unmyelinated axons. Anesthesiology 1984, 60, 111–120. [Google Scholar] [CrossRef] [PubMed]
  328. Brodin, P. Differential inhibition of A, B and C fibres in the rat vagus nerve by lidocaine, eugenol and formaldehyde. Arch. Oral Biol. 1985, 30, 477–480. [Google Scholar] [CrossRef] [PubMed]
  329. Nakamura, M.; Jang, I.-S. Indomethacin inhibits tetrodotoxin-resistant Na+ channels at acidic pH in rat nociceptive neurons. Neuropharmacology 2016, 105, 454–462. [Google Scholar] [CrossRef] [PubMed]
  330. Joshi, S.K.; Mikusa, J.P.; Hernandez, G.; Baker, S.; Shieh, C.C.; Neelands, T.; Zhang, X.F.; Niforatos, W.; Kage, K.; Han, P.; et al. Involvement of the TTX-resistant sodium channel Nav 1.8 in inflammatory and neuropathic, but not post-operative, pain states. Pain 2006, 123, 75–82. [Google Scholar] [CrossRef] [PubMed]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kumamoto, E. Inhibitory Actions of Clinical Analgesics, Analgesic Adjuvants, and Plant-Derived Analgesics on Nerve Action Potential Conduction. Encyclopedia 2022, 2, 1902-1934. https://doi.org/10.3390/encyclopedia2040132

AMA Style

Kumamoto E. Inhibitory Actions of Clinical Analgesics, Analgesic Adjuvants, and Plant-Derived Analgesics on Nerve Action Potential Conduction. Encyclopedia. 2022; 2(4):1902-1934. https://doi.org/10.3390/encyclopedia2040132

Chicago/Turabian Style

Kumamoto, Eiichi. 2022. "Inhibitory Actions of Clinical Analgesics, Analgesic Adjuvants, and Plant-Derived Analgesics on Nerve Action Potential Conduction" Encyclopedia 2, no. 4: 1902-1934. https://doi.org/10.3390/encyclopedia2040132

Article Metrics

Back to TopTop