Actions of Analgesics on Nerve Conduction: Comparison
Please note this is a comparison between Version 5 by Peter Tang and Version 6 by Peter Tang.

The aAction potential (AP) conduction in nerve fibers plays a  crucial role in transmitting nociceptive information from the periphery to the cerebral cortex. NIt is possible that nerve AP conduction inhibition possibly results in analgesia. It is Many of analgesics are well-known that many analgesics so suppress nerve AP conduction and voltage-dependent sodiumgated Na+ and potassiumK+ channels that are involved in producing APs. The c AP conduction. 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 ne measure to know whether nerve AP conduction inhibition may contribute to at least a part of their as affected by analgesic effects. 

  • antinociception
  • analgesic
  • analgesic adjuvant
  • plant-derived compound
  • nerve conduction
  • sciatic nerve
  • compound action potential
  • sodium channel
  • potassium channel
  • Na+ channel
  • K+ channel
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. Introduction

Information of nociceptive stimuli given to the periphery is mainly transmitted through primary-afferent myelinated Aδ and unmyelinated C fibers to the spinal cord and brain stem and then to the brain by action potential (AP) conduction in nerve fibers and chemical transmission at synapses (see

[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

for review). Acute nociceptive pain caused by tissue injury or damage is a physiological mechanism that serves to protect a person against injury and 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 persist or recur for longer than three months, is a debilitating disease accompanied by spontaneous pain, etc. and is often resistant to analgesics such as NSAIDs and opioids. One of the chronic pain, neuropathic pain, occurs as a result of direct injury of the peripheral nervous system (PNS) and damage to the central nervous system (CNS), and is characterized by a hyper-excitability of neurons near injured or damaged neuronal tissues (see

[5]. This type of pain is alleviated by using analgesic adjuvants such as local anesthetics, antiepileptics, antidepressants, and α

for review). This type of pain is alleviated by using analgesic adjuvants such as local anesthetics, antiepileptics, antidepressants and α

2-adrenoceptor agonists

-adrenoceptor agonists (see

[6][7][8][9][10][11][12][13][14][15]. Although analgesics and analgesic adjuvants generally depress excitatory synaptic transmission

for review). Although analgesics and analgesic adjuvants generally depress excitatory synaptic transmission (see

[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

for review), many of their drugs possibly suppress nerve AP conduction, in part contributing 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 (see

[19][20][21].

for review).

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

AP conduction is mediated by voltage-gated Na+ and K+ channels expressed in nerve fibers. Thus, a depolarizing stimulus given to a point of nerve fiber activates Na+ channels located in membranes of the fiber, leading to Na+ entry to the cytoplasm, owing to the gradient of the electrochemical potential of Na+, resulting in a self-regenerative production of AP. This in turn produces an outward current (membrane depolarization) in a fiber point adjacent to the point to open other Na+ channels. Such an AP production subsides by a subsequent Na+ channel inactivation and K+ channel activation (see

[22][23].

for review).

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

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. Voltage-gated Na+-channel blocker tetrodotoxin (TTX)-sensitive and fast-conducting (possibly primary-afferent Aα fiber-mediated) CAPs can be easily observed in the sciatic nerve trunk isolated from frogs by exposing the nerve trunk to air (known as air-gap method). A half-peak duration of the CAP was increased by a voltage-gated delayed-rectifier K+-channel inhibitor tetraethylammonium with no change in its peak amplitude, indicating an involvement of K+ channels (see

[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

for review). Although the frog sciatic nerve exhibits not only fast-conducting but also slow-conducting (Aδ-fiber and C-fiber mediated) CAPs, the latter CAP has much smaller peak amplitude and conduction velocity than the former CAP

[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

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 such as NSAIDs

[26], many types of opioids such as tramadol

, many kinds of opioids including tramadol [27][28], many amide- and ester-type local anesthetics

[27][28], many amide- and ester-type local anesthetics [29], antiepileptics [30], antidepressants

[31][32], antiepileptics [29], antidepressants [33], an α2-adrenoceptor agonist dexmedetomidine (DEX; (+)-(S)-4-[1-(2,3-dimethylphenyl)-ethyl]-1H-imidazole; [34]), and many kinds of antinociceptive plant-derived compounds (see [35] for review).

2. Actions of Analgesics on Nerve AP Conduction

2.1. NSAIDs

NSAIDs produce antinociception by various mechanisms including inhibition of the synthesis of prostaglandins from arachidonic acid by inhibiting the cyclooxygenase enzyme ([36][37]; see [38][39][40][41] for review), activation of several K+ channels ([42][43][44][45][46]; see [47][48] for review), inhibition of acid-sensitive ion channels [49] and transient receptor potential (TRP) channels [50][51], depletion of substance P [52], an interaction with the adrenergic nervous system [53] and an involvement of opioids [54][55] and endocannabinoids (see [56] for review). The idea that antinociception is produced by mechanisms other than cyclooxygenase inhibition is supported by the observation that there is a dissociation between antinociception and anti-inflammation produced by NSAIDs [57].

An acetic acid-based NSAID diclofenac inhibited frog sciatic nerve CAPs in a partially reversible manner with a half-maximal inhibitory concentration (IC50) value of 0.94 mM in a concentration range of 0.01-1 mM. Another acetic acid-based NSAID aceclofenac, which is a carboxymethyl ester of diclofenac, also depressed CAPs in a concentration range of 0.01-1 mM with an IC50 of 0.47 mM, a value smaller than that of diclofenac. Other acetic acid-based NSAIDs exhibited a similar CAP inhibition, albeit being smaller in extent than diclofenac and aceclofenac. Indomethacin at 1 mM reduced CAP peak amplitudes by 38%, and acemetacin, where the -OH group of indomethacin is substituted by -OCH2COOH, at 0.5 mM reduced CAP peak amplitudes by 38%. Etodolac at 1 mM reduced CAP peak amplitudes by only 15%, and sulindac and felbinac at 1 mM unaffected CAP peak amplitudes [26].

Frog sciatic nerve CAP peak amplitudes were reduced by fenamic acid-based NSAIDs (tolfenamic acid, meclofenamic acid, mefenamic acid and flufenamic acid) having a chemical structure similar to those of diclofenac and aceclofenac in a concentration-dependent manner. Tolfenamic acid reduced CAP peak amplitudes with an IC50 value of 0.29 mM in a concentration range of 0.01-0.2 mM. Meclofenamic acid, where the chloro group bound to the benzene ring of tolfenamic acid is altered in number and position, had an IC50 value of 0.19 mM in a concentration range of 0.01-0.5 mM. Moreover, mefenamic acid, where the chloro group bound to the benzene ring of tolfenamic acid is replaced by methyl group, reduced CAP peak amplitudes in a concentration range of 0.01-0.2 mM (by 16% at 0.2 mM). Flufenamic acid, where one out of two methyl groups bound to the benzene ring of mefenamic acid is lacking and another one is replaced by -CF3, exhibited an IC50 of 0.22 mM, a value similar to those of tolfenamic acid and meclofenamic acid [26]. With respect to other types of NSAIDs, frog sciatic nerve CAPs were not affected by salicylic acid-based (aspirin; 1 mM), propionic acid-based (ketoprofen, naproxen, ibuprofen, loxoprofen and flurbiprofen; each 1 mM) and enolic acid-based NSAIDs [meloxicam (0.5 mM) and piroxicam (1 mM)] [26].

On the other hand, frog sciatic nerve CAPs were concentration-dependently inhibited by 2,6-dichlorodiphenylamine and N-phenylanthranilic acid, which are not NSAIDs while being similar in chemical structure to diclofenac and tolfenamic acid. Thus, 2,6-dichlorodiphenylamine lacks the -CH2COOH group of diclofenac and N-phenylanthranilic acid lacks chloro and methyl groups bound to the benzene ring of tolfenamic acid. 2,6-Dichlorodiphenylamine activity was seen in a concentration range of 0.001-0.1 mM with the extent of 45% at 0.1 mM. N-phenylanthranilic acid activity was seen in a concentration range of 0.01-2 mM with the extent of 23% at 1 mM [26].

The NSAIDs-induced CAP inhibition would be mediated by a suppression of TTX-sensitive voltage-gated Na+ channels that are involved in producing frog CAPs. Consistent with this idea, diclofenac reduced TTX-sensitive Na+-channel current amplitudes in rat dorsal root ganglion (DRG) [58] and mouse trigeminal ganglion neurons [59]. A similar Na+-channel inhibition produced by diclofenac has been reported in rat myoblasts [60] and ventricular cardiomyocytes [61]. Like diclofenac, flufenamic acid reduced Na+-channel current amplitudes in rat hippocampal CA1 neurons [62][63][64]. Although IC50 value (0.22 mM) for flufenamic acid to suppress frog sciatic nerve CAPs was close to that (0.189 mM) of Na+ channel inhibition in rat hippocampal CA1 neurons [64], IC50 value (0.94 mM) for diclofenac-induced CAP inhibition was much larger than those (0.014 and 0.00851 mM in rat DRG neurons and myoblasts, respectively) for Na+ channel inhibition [58][60]. Rank order for CAP inhibition by NSAIDs at 0.5 mM was flufenamic acid > diclofenac > indomethacin >> aspirin = naproxen = ibuprofen [26].   This order was partly similar to those for Na+ channel suppression in rat DRG neurons (diclofenac > flufenamic acid > indomethacin > aspirin; [58]) and also in rat cardiomyocytes (diclofenac > naproxen ≥ ibuprofen; [61]). Diclofenac at 0.3 mM reduced by about 20% TTX-resistant Na+-channel current amplitudes in rat trigeminal ganglion neurons [65]. TTX-resistant Nav1.8 channel current amplitudes were reduced by flufenamic acid and tolfenamic acid at 0.1 mM with the extents of about 30 and 30%, respectively [66]. TTX-sensitive Nav1.7 channel currents were more sensitive to flufenamic acid and tolfenamic acid (extent at 0.1 mM: about 60 and 70%, respectively) than TTX-resistant Nav1.8 ones [66]. Alternatively, NSAIDs inhibited the extent of chemical irritation-induced activity increase of cat corneal sensory nerve fibers; this inhibition was different in extent among different types of NSAIDs [59][67]. NSAIDs-induced Na+-channel inhibition appeared to be distinct in magnitude among preparations. Concentrations needed for NSAIDs to significantly inhibit frog sciatic nerve CAPs were generally higher than those necessary for Na+ channel inhibition. This result may be due to various factors such as the involvement of not only Na+ channels but also K+ channels in CAP peak amplitudes. As far as I know, it has not been reported how the aceclofenac, indomethacin, etodolac, acemetacin, meclofenamic acid and mefenamic acid affect voltage-gated Na+ channels. Table 1 summarizes the actions of NSAIDs on frog sciatic nerve fast-conducting CAPs together with their IC50 values (see also [68]).

More effective NSAIDs in inhibiting frog sciatic nerve CAPs have two benzene rings that bind a hydrophilic substituent group, both of which rings are linked by -NH- (see Figs. 1Aa, 1Ba, 3Aa, 3Ba, 3Da 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 (see Section 3.1) but not in the other NSAIDs [26]. Mefenamic acid, where one of the two benzene rings has a hydrophobic substituent group (see Fig. 3Ca in [26]), seemed to be less effective in CAP inhibition, albeit not tested at a higher concentration owing to a less solubility of this drug (see above). CAPs were effectively inhibited by 2,6-dichlorodiphenylamine and N-phenylanthranilic acid, which are similar in chemical structure to NSAIDs having two benzene rings (see Figs. 4Aa and 4Ba in [26]), albeit they are not NSAIDs. CAPs were also suppressed by the endocrine disruptor bisphenol A having two benzene rings that bind a hydrophilic group such as -OH

[31], dexmedetomidine (DEX; (+)-(S)-4-[1-(2,3-dimethylphenyl)-ethyl]-1H-imidazole, which is an α

.

There is much evidence showing that the other actions of NSAIDs are dependent on their chemical structures. For example, an involvement of NO-cyclic GMP-K+ channels in NSAIDs-induced antinociception depended on their chemical structures [45][69]. Nonselective cation channels in the rat exocrine pancreas were inhibited by flufenamic acid and mefenamic acid but not indomethacin, aspirin and ibuprofen [70]. There was a difference between diclofenac and aceclofenac in inhibiting TRP melastatin-3 channels [51]. Although TRP ankyrin-1 (TRPA1) channels were depressed or activated by NSAIDs, such an activity also differed in extent among NSAIDs [71]. Moreover, a difference was seen among NSAIDs in the activities of electron transport system or mitochondrial oxidative phosphorylation that may produce NSAIDs’ adverse side effects [72].

As stated above, the concentrations for NSAIDs to inhibit frog sciatic nerve CAPs are generally much higher than those for voltage-gated Na+-channel inhibition. Such high concentrations will be possible when NSAIDs are used in the direct vicinity of nerve fibers. At least a part of analgesia produced by NSAIDs used as a dermatological drug may be explained by a nerve AP conduction inhibition through suppressed voltage-gated Na+ channels [73].

2.2. Opioids

Opioids suppress glutamatergic excitatory transmission by activating opioid receptors expressed in the CNS including the central terminals of primary-afferent fibers, leading to antinociception ([74][75][76]; see [77][78] for review). Opioid receptors are located in not only central but also peripheral terminals of primary-afferent neurons; peripheral terminal opioid receptors are known to be involved in antinociception ([79][80][81][82][83]; see [84] for review). 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 [85], AP conduction in peripheral nerve fibers is generally inhibited by opioids. For example, opioids such as fentanyl and sufentanil reduced the peak amplitudes of CAPs recorded from peripheral nerve fibers [86] and depressed peripheral nerve AP conduction [87]. A morphine-induced CAP inhibition in mammalian peripheral nerve fibers was sensitive to a nonspecific opioid-receptor antagonist naloxone, indicating an involvement of opioid receptors [88]. Consistent with this observation, binding and immunohistochemical studies have demonstrated the localization of opioid receptors in mammalian peripheral nerve fibers [89][90][91].

2.2.1. Tramadol

Tramadol [(1RS; 2RS)-2-[(dimethylamino) methyl]-1-(3-methoxyphenyl)-cyclohexanol hydrochloride] is an orally-active and clinically-used opioid in the CNS [92]. In animals and humans, tramadol is metabolized to various compounds including mono-O-desmethyl-tramadol (M1) through N- and O-demethylation [93]; M1 is a therapeutically active drug to alleviate pain [92]. Among cellular mechanisms for the tramadol’s antinociceptive effect, there is μ-opioid receptor activation [94][95]. This idea is supported by the highest affinity of M1 among the metabolites of tramadol for cloned μ-opioid receptors. M1 inhibited glutamatergic excitatory transmission in spinal lamina II neurons which play a pivotal role in regulating nociceptive transmission to the spinal dorsal horn from the periphery, resulting in reducing the excitability of the neurons [96][97][98]. In addition to such a central activity, tramadol has a local anesthetic effect following its intradermal injection in patients ([99][100][101]; see [102] for review). This result was consistent with in vivo studies showing an inhibition of a spinal somatosensory evoked potential, produced by a direct application of tramadol to the rat sciatic nerve [103]. Tramadol used as an adjuvant to local anesthetics has been reported to prolong the duration of sensory block and analgesia [104].

CAPs recorded from the frog sciatic nerve were reduced in peak amplitude by tramadol in a concentration-dependent manner in a range of 0.2 to 5 mM [27]. A similar tramadol’s CAP inhibitory action has been demonstrated by other investigators in the frog [105] and rat sciatic nerve [106][107]. According to the analysis based on the Hill equation, IC50 value for tramadol to reduce frog sciatic nerve CAP amplitude was 2.3 mM, a value being smaller by about three-fold than that (6.6 mM) reported by Mert et al. [105] for the frog sciatic nerve. Tramadol also suppressed rat sciatic nerve CAPs (37% peak amplitude reduction at 4 mM; [106]) with an extent less than that obtained by Katsuki et al. [27] for frog sciatic nerve CAPs. The tramadol action in the frog sciatic nerve was unaffected by the pretreatment of the nerve with naloxone (0.01 mM) and a μ-opioid receptor agonist (D-Ala2, N-Me-Phe4, Gly5-ol)enkephalin (DAMGO; 1 μM) had no effect on frog sciatic nerve CAPs [27]. As different from tramadol, M1 inhibited CAPs by much smaller extents (see below), although M1 has a higher affinity for μ-opioid receptors than tramadol, the chemical structure of which is similar to that of M1 [108]. These results indicate that opioid receptors are not involved in tramadol-induced CAP inhibition [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 [103]. Jaffe and Rowe [87] also have reported a naloxone-resistant nerve AP conduction inhibition produced by opioids.

Although tramadol inhibits noradrenaline (NA) and serotonin (5-hydroxytryptamine; 5-HT) reuptake at concentrations enough to activate μ-opioid receptors [109][110], a combination of NA and 5-HT reuptake inhibitors (desipramine and fluoxetine, respectively; each 10 μM; see Section 3.3) unaffected frog sciatic nerve CAPs, indicating that CAP inhibition was not mediated by NA and 5-HT reuptake inhibition [27].

It is possible that the tramadol-induced CAP inhibition is mediated by an inhibition of voltage-gated Na+ and K+ channels involved in AP production. Tramadol concentration-dependently reduced TTX-sensitive Na+ channel current amplitude with an IC50 value of 0.194 mM in DRG neuroblastoma hybridoma cell line ND7/23 cells [111] and also with an IC50 value of 0.103 mM in HEK293 cells expressing rat TTX-sensitive Nav1.2 channels [112]. These values were smaller than that (2.3 mM) of IC50 for frog sciatic nerve CAP amplitude reduction [27]. Tramadol also decreased the current amplitude of delayed rectifier K+-channels (Kv3.1a type) expressed in NG 108-15 cells with an IC50 of 0.025 mM; this value was much less than 2.3 mM [113]. Such IC50 values of tramadol for CAP, Na+ and K+ channel inhibition were higher than its clinically relevant concentration of about 2 μM in serum [98][114].

On the other hand, frog sciatic nerve CAPs were unaffected by M1 (1-2 mM), as different from tramadol. Moreover, in the frog sciatic nerve exhibiting CAP inhibition by tramadol (1 mM; [27]), M1 at 5 mM reduced CAP peak amplitudes by 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 dorsal root-evoked excitatory postsynaptic currents was investigated by applying the patch-clamp technique to lamina II neurons in spinal cord slices [97]. Interestingly, tramadol has -OCH3 while M1 has -OH bound to the benzene ring, and thus the methyl group is present in tramadol but not M1 (see Fig. 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. 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 (which has -OCH3 in place of -OH in morphine) and ethylmorphine (where -OH of morphine is replaced by -OCH2CH3; see Fig. 7A in [28] for their chemical structures). CAP peak amplitude was reduced by morphine in a concentration-dependent manner with an extent of 15% at 5 mM. Codeine at 5 mM reduced CAP peak amplitude by 30%. Ethylmorphine inhibited CAPs more effectively than morphine and coceine with an IC50 value of 4.6 mM (inhibition at 5 mM: 61%). The activities of morphine, codeine and ethylmorphine were resistant to naloxone (0.01 mM). Naloxone at 1 mM by itself reduced by 9% CAP peak amplitudes, while unaffecting 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 [86][87][115]. On the contrary, Hunter and Frank [116] have reported a naloxone-sensitive CAP inhibition in the frog sciatic nerve. A sequence of the CAP peak amplitude reduction produced by opioids was ethylmorphine > codeine > morphine, indicating that CAP inhibition increases in extent with an increase in the number of -CH2. This result was consistent with the relationship between tramadol and M1, as mentioned above. It is of interest to note that morphine, codeine and ethylmorphine are quite distinct in chemical structure from tramadol and M1 (see [117] for review). Since the increase in -CH

2-adrenoceptor agonist;

number enhances lipophilicity of opioids, it is suggested that lipophilic opioid-channel interaction plays a crucial role in inhibiting nerve AP conduction, as shown for local anesthetics [118][119]. Consistent with this idea, a potency of rat sciatic nerve CAP inhibition was in the order of isopropylcocaine > cocaethylene > cocaine [120]. Interestingly, the sequence of an affinity of opioids for μ-opioid receptors is morphine > codeine > ethylmorphine [121]; this order is reversed to that of CAP inhibition. This result supports the idea that frog sciatic nerve CAP inhibition produced by opioids is not due to opioid receptor activation.

CAP inhibition similar to that in the frog sciatic nerve has been reported in the mammalian peripheral nerve, although its extent is different among preparations. Frog sciatic nerve CAP peak amplitude reduction (about 30%) produced by codeine (5 mM) was much smaller than that (about 70%) in the rat phrenic nerve, while there was not so a large difference in morphine (5 mM) action (about 10%) between the two preparations. Morphine sensitivity was less in the frog sciatic nerve than rabbit and guinea-pig vagus nerves whose CAP peak amplitudes were reduced by 20-32% at 0.5 mM [88]. Intracellularly-recorded APs in rat DRG neurons having Aα/β myelinated primary-afferent fibers were also inhibited by opioids with a sequence of ethylmorphine > codeine ≥ morphine (IC50 = 0.70, 2.5 and 2.9 mM, respectively) in AP peak amplitude reduction. These AP inhibitions were also resistant to naloxone (0.01 mM) [122].

AP conduction in peripheral nerve fibers is inhibited by many drugs including narcotics, antiepileptics, local anesthetics, alcohols and barbiturates, suggesting that the drugs may interact with membrane bilayers in a nonspecific manner [123]. However, the above-mentioned chemical structure-specific CAP inhibition produced by opioids indicates that opioids act on membrane proteins such as voltage-gated Na+ and K+ channels (see [124] for review). In support of this idea, morphine inhibited peak Na+ channel currents and steady-state K+ channel currents recorded from frog sciatic nerve single myelinated nerve fibers, resulting in the prolongation of APs [125]. The intracellular application of morphine resulted in the reduction of voltage-gated Na+ and K+ channel current amplitudes in squid giant axons [126]. Bath application of morphine led to the reduction of TTX-sensitive Na+ channel current amplitude in DRG neuroblastoma hybridoma cell line ND7/23 cells with an IC50 value of 0.378 mM [111], whereas TTX-sensitive Nav1.2 channels located in HEK293 cells were unaffected by morphine at 1 mM [112]. Supporting the idea about ion channel inhibition, an opioid meperidine, which was used for AP conduction blockade and thus analgesia, depressed Na+-channels in a manner similar to that of lidocaine [127]. Table 1 summarizes the inhibitory actions of opioids on frog sciatic nerve fast-conducting CAPs together with their IC50 values (see also [68]).

In clinical practice, many of pain treatments by using opioids are due to systemic administration of centrally-penetrating opioids, leading to their actions in the PNS and CNS, both of which contribute to analgesia (see [128] for review). Administration of opioids into the nerve sheath also could alleviate pain (for instance, see [129]). It is likely that centrally-administrated opioids act on not only the CNS but also the PNS, because opioids are transported to the periphery from brain by P-glycoprotein [130]. Supporting this idea, subcutaneous administration of blood brain barrier-impermeable N-methyl-morphine produced antinociception in an acetic acid-writhing model in mice [79]. A subcutaneously-administrated opioid loperamide, which is impermeable into the brain, had an antinociceptive effect in the formalin test in rats [81]. Thus, nerve AP conduction inhibition produced by opioids might contribute to local analgesia following the peripheral perineural administration of opioids (for instance, see [131]) that may lead to a direct action of opioids at high doses on peripheral nerves. Peripherally-applied codeine might have a similar effect to that of morphine, because codeine is metabolized to morphine via O-demethylation in humans and animals ([132][133]; see [128] for review).

3. Actions of Analgesic Adjuvants on Nerve AP Conduction

3.1. Local Anesthetics

Local anesthetics inhibit both voltage-gated Na+ and K+ channels ([118]; see [124][134][135] for review). Owing to this inhibition, local anesthetics have been used to alleviate neuropathic pain in the hope of suppressing nerve AP conduction in animals [136][137] and humans [138][139][140][141], 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 (see [134] for review]. Various types of local anesthetics are reported to activate TRPA1 channels in the central terminals of primary-afferent neurons in lamina II of the rat spinal dorsal horn [142][143], and also TRPA1 and TRP vanilloid-1 (TRPV1) channels in rodent DRG neurons [144][145].

3.1.1. Amide-type Local Anesthetics

Frog sciatic nerve CAPs were reversibly reduced in peak amplitude by an amide-type local anesthetic lidocaine, which blocks nerve AP conduction [105][106][107][135], in a concentration range of 0.1 to 2 mM with an IC50 value of 0.74 mM [28]. This IC50 value was somewhat larger than that (0.204 mM) for voltage-gated Na+-channel current amplitude reduction while being smaller than that (1.118 mM) for voltage-gated K+-channel current amplitude reduction in Xenopus laevis sciatic nerve fibers [119]. Rat TTX-resistant Na+ channel current amplitude was reduced by lidocaine with an IC50 of 0.073 mM [146], a value 10-fold lower than that for frog sciatic nerve CAP amplitude reduction. At a least of antinociception produced by systemically-applied lidocaine in humans [147] may be attributed to its inhibitory effect on nerve AP conduction.

A similar reversible CAP amplitude reduction was produced by another amide-type local anesthetic ropivacaine, which exhibits a longer duration of action in terms of nerve AP conduction block than lidocaine does ([148]; see [149]] for review); this reduction was concentration-dependent in a range of 0.01-1 mM with an IC50 value of 0.34 mM [27]. The frog sciatic nerve ropivacaine-induced CAP amplitude reduction was almost comparable in extent to that (about 30% at 0.2 mM) in rabbit vagus nerve A fibers [150]. Frog sciatic nerve IC50 values for lidocaine and ropivacaine (0.74 and 0.34 mM, respectively) were not so distinct from those (0.28 mM for both lidocaine and ropivacaine) for fast-conducting CAP amplitude reduction in the rat sciatic nerve [151]. Moreover, an amide-type local anesthetic prilocaine also reversibly reduced frog sciatic nerve CAP peak amplitudes in a concentration-dependent manner in a range of 0.01-5 mM with an IC50 value of 1.8 mM [31].

As amide-type local anesthetics, there are levobupivacaine and its racemic bupivacaine, the former of which has a lower risk of cardiovascular and CNS toxicity than the latter ([152]; see [153] for review). Frog sciatic nerve CAP peak amplitudes were reversibly reduced by levobupivacaine in a concentration-dependent manner in a range of 0.05-1 mM with an IC50 value of 0.23 mM [29]. This IC50 value was close to that (0.22 mM) reported previously for a tonic levobupivacaine-induced inhibition of frog sciatic nerve CAPs [152] and to that (0.264 mM) for a tonic levobupivacaine-induced suppression of voltage-gated Na+-channel currents recorded at -100 mV in GH-3 neuroendocrine cells [154]. As shown previously [152], the levobupivacaine-induced CAP amplitude reduction in the frog sciatic nerve was smaller than that of bupivacaine (their extents at 0.5 mM: 45 and 76%, respectively; [29]). This frog sciatic nerve bupivacaine activity was smaller than that (IC50 = 0.027 mM) for Na+-channels in Xenopus laevis sciatic nerve fibers [119], that (IC50 = 0.178 mM) for TTX-sensitive Na+ channels in DRG neuroblastoma hybridoma cell line ND7/23 cells [155] and that (IC50 = 0.190 mM) for Na+ channels in rat clonal pituitary GH3 cells [156]. Voltage-gated K+ channels in Xenopus laevis sciatic nerve fibers were also inhibited by bupivacaine with a sensitivity (IC50 = 0.092 mM) less than that for Na+ channels [119].

3.1.2. Ester-type Local Anesthetics

As a classic ester-type local anesthetic, there is a compound derived from the coca plant Erythroxylon coca, cocaine, which is well-known to inhibit nerve AP conduction ([115][157][158]; see [159] for review). Frog sciatic nerve CAP peak amplitude was reversibly reduced by cocaine in a concentration-dependent manner in a range of 0.01-2 mM with an IC50 of 0.80 mM [28], a value similar to that (0.74 mM) of lidocaine in the frog sciatic nerve [27]. The cocaine’s IC50 value was about 4-fold larger than that (about 0.2 mM) in the rat phrenic nerve [115]. Although cocaine (40 μM) reduced mouse phrenic nerve CAP peak amplitudes by 26% [158], such a reduction in the frog sciatic nerve was produced at a concentration of about 300 μM [28]. There was an almost similar CAP amplitude reduction by cocaine in the frog and rat sciatic nerve (frog: 30% at 0.5 mM; rat: 40% at 0.375 mM; see [120]). There is much evidence for an inhibition by cocaine of voltage-gated Na+ channels (for example, see [120][160][161]); a cocaine (0.05 mM)-induced tonic (TTX-resistant) Nav1.5 channel current amplitude reduction is about 70% [161]. Cocaine and lidocaine suppressed voltage-gated Na+ channels in a competitive manner [162]. Cocaine also inhibited delayed rectifier K+ channels in central snail neurons [163].

Another ester-type local anesthetic procaine [164] also reversibly reduced frog sciatic nerve CAP peak amplitudes in a concentration-dependent manner in a range of 0.1-5 mM with an IC50 value of 2.2 mM

[32]), and diverse kinds of antinociceptive compounds isolated from plants

. This IC50 value was close to those (2-5 mM) obtained by other researchers [165][166] in the same preparation and also to that (about 1 mM) reported in the rat sciatic nerve [165]. Moreover, a ratio (2.2 mM/0.74 mM) of the procaine’s IC50 value to that of lidocaine [27] in reducing frog sciatic nerve CAP amplitudes was comparable to a ratio (0.53%/0.14%) of procaine concentration, necessary to block motor nerve AP conduction by 50%, to lidocaine’s one in rats [167]. On the other hand, procaine activity in the frog sciatic nerve was 37-fold smaller than that (IC50 = 0.060 mM) in reducing voltage-gated Na+-channel current amplitude in Xenopus laevis sciatic nerve fibers [119]. Procaine also suppressed voltage-gated K+ channels in this preparation with an IC50 value (6.303 mM) larger than that for Na+ channels [119].

There is an ester-type local anesthetic benzocaine (ethyl 4-aminobenzoate) that is used for not only topical anesthesia in clinical medicine (see [168] for review) but also amphibian anesthesia ([169]; see [170][171] for review). Frog sciatic nerve CAP peak amplitudes were reversibly reduced by benzocaine in a concentration-dependent manner in a range of 0.01-2 mM with an IC50 value of 0.80 mM (73% reduction at 1 mM; [30]). The rat sciatic nerve exhibited a similar benzocaine-induced CAP inhibition (37% inhibition at 1.3 mM; [106]). The benzocaine activity was similar to those of cocaine and lidocaine.

Another ester-type local anesthetic tetracaine also reduced frog CAP peak amplitudes in a reversible and concentration-dependent manner with an IC50 value of 0.014 mM [34]. This IC50 value is not so different from that (0.0063 mM) of frog sciatic nerve fibers, as reported by Starke at al. [172], and also from that (0.009 mM) obtained for rabbit A nerve fibers [173]. On the other hand, the tetracaine activity in the frog sciatic nerve was 19-fold smaller than that (IC50 = 0.0007 mM) for voltage-gated Na+-channel current amplitude reduction in Xenopus laevis sciatic nerve fibers [119]. Voltage-gated K+-channel current amplitudes in this preparation were also reduced by tetracaine with an IC50 value (0.946 mM) much larger than that for Na+ channels [119]. Tetracaine had much more effectiveness than procaine and also lidocaine and bupivacaine in both frog CAP and Xenopus laevis Na+-channel current inhibition.

Frog sciatic nerve CAP peak amplitudes were also reduced by a non-amide- and non-ester-type local anesthetic pramoxine. This inhibitory action of pramoxine was concentration-dependent in a range of 0.001-1 mM with an IC50 value of 0.21 mM and subsided with a slow time course after its washout [31]. Table 1 summarizes the inhibitory actions of local anesthetics on frog sciatic nerve fast-conducting CAPs together with their IC50 values (see also [68]).

3.2. Antiepileptics

Antiepileptics have various actions such as glutamate-receptor inhibition, GABAA-receptor activation, voltage-gated Na+- and Ca2+-channel inhibition (see [13][174] for review). Antiepileptics are well-known to inhibit neuropathic pain (for example, see [175]). As indicated by the action on Na+ channels, it is possible that the neuropathic pain alleviation is due to nerve AP conduction inhibition.

Frog sciatic nerve CAP peak amplitudes were reduced by a phenyltriazine derivative (lamotrigine; 3,5-diamino-6-(2,3-dichlorophenyl)-1,2,4-triazine) that suppressed voltage-gated Na+ channels [176] and also attenuated central post-stroke pain and painful diabetic poly-neuropathy [7]. This sciatic nerve lamotrigine activity was partially reversible and concentration-dependent in a range of 0.02-0.5 mM with an IC50 value of 0.44 mM [29]. This value was similar to IC50 value (0.641 mM at -90 mV) for lamotrigine to inhibit TTX-sensitive human brain type IIA Na+ channels expressed in Chinese hamster ovary cells [176]. A similar CAP amplitude reduction was seen by an iminostilbene derivative carbamazepine (5H-dibenz[b,f]azepine-5-carboxamide; [29]), which is distinct in chemical structure from lamotrigine while inhibiting voltage-gated Na+ channels [177]. Carbamazepine is reportedly effective in attenuating trigeminal neuralgia (see [178][179] for review). As distinct from lamotrigine’s one, the carbamazepine-induced CAP inhibition in the frog sciatic nerve was completely reversible. This carbamazepine activity was concentration-dependent in a range of 0.05-1 mM with an IC50 value of 0.50 mM [29]. Carbamazepine and lamotrigine reduced Na+-channel current amplitudes in N4TG1 mouse neuroblastoma cells with IC50 values similar to each other [180], an observation being consistent with the fact that the two antiepileptics had comparable IC50 values in reducing frog sciatic nerve CAP amplitudes.

Oxcarbazepine (10,11-dihydro-10-oxo-5H-dibenz[b,f]azepine-5-carboxamide; [181]), where there is a keto substitution at the 10,11 position of the dibenzazepine nucleus of carbamazepine, reduced frog sciatic nerve CAP peak amplitudes with an efficacy less than that of carbamazepine [29]. Oxcarbazepine is known to be effective in attenuating painful diabetic neuropathy [7] and trigeminal neuralgia [178]. Oxcarbazepine activity was partially reversible and concentration-dependent in a range of 0.02-0.7 mM. CAP amplitude reduction (40%) by oxcarbazepine (0.7 mM) was somewhat smaller in extent than that (57%) of carbamazepine (0.7 mM). Each of lamotrigine, carbamazepine and oxcarbazepine at 0.5 mM increased a threshold to evoke frog sciatic nerve CAPs [29]. Consistent with this observation, their antiepileptics reduced voltage-gated Na+-channel current amplitudes with a shift of 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%; [29]) was much smaller in extent than that for TTX-sensitive Na+-channel current inhibition 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 seizures produced by maximal electroshock in rats [181].

Another antiepileptic phenytoin (hydantoin derivative, 5,5-diphenylhydantoin; which suppresses voltage-gated Na+ channels [184] and attenuates paroxysm in trigeminal neuralgia [179]) inhibited frog sciatic nerve CAPs with a small extent in a concentration-dependent manner in a range of 0.01-0.1 mM; this extent was only 15% at 0.1 mM [29]. The frog sciatic nerve phenytoin activity was less than those of rat cortical and human type IIA Na+ channels (60-90% amplitude reduction by phenytoin at 0.1 mM at -60 mV) [176][184]. As distinct from frog sciatic nerve CAPs, voltage-gated Na+ channel current amplitudes were reduced 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 of Na+ channels in rat hippocampal CA1 neurons [183]. A sensitivity of voltage-gated Na+ channels to phenytoin appeared to be distinct in extent among distinct types of the channel. Supporting this idea, phenytoin activities were distinct in magnitude among human Nav1.1, Nav1.2, Nav1.3 and Nav1.4 α-subunits (all of which are TTX-sensitive) expressed in HEK293 cells [185]. Moreover, there was a difference in the properties and accessibilities of Na+ channels between frog and rat myelinated nerves [186].

Antiepileptics having an ability to suppress CAPs are similar in chemical structure to NSAIDs in that lamotrigine, carbamazepine and oxcarbazepine have two unsaturated six-membered rings (see Figs. 1a, 2aA and 2bA in [29] for the chemical structures of the three antiepileptics). Carbamazepine and diclofenac appear to have a common or closely-related binding site because of an occlusion of their effects on voltage-gated Na+ channels [63].

On the other hand, frog sciatic nerve CAPs were not affected by other antiepileptics, gabapentin (1-(aminomethyl)cyclohexaneacetic acid; which is related to GABA in chemical structure and attenuates post-herpetic neuralgia [179]]), topiramate (2,3:4,5-bis-O-(1-methylethylidene)-β-D-fructopyranose sulfamate; which alleviates various neuropathic pains such as intercostal neuralgia and trigeminal neuralgia [13]) and sodium valproate (2-propylpentanoic acid sodium salt; which relieves diabetic neuropathic pain [13]), at a high concentration such as 10 mM [29]. The less effectiveness of gabapentin and sodium valproate in the frog sciatic nerve was similar to that for human type IIA Na+ channels [176]. The human Na+ channels were hardly affected by gabapentin at concentrations of less than 3 mM [176]. Gabapentin’s antinociceptive action would be due to its binding to the α2δ-1 subunit of voltage-gated Ca2+ channels, resulting in inhibited Ca2+ entry in nerve terminals which in turn attenuates the release of neurotransmitters from there (see [187] for review); gabapentin also may interrupt an interaction between N-methyl-D-aspartate (NMDA)-receptor (a subtype of glutamate receptors) channels and the α2δ-1 subunit of voltage-gated Ca2+ channels in postsynaptic neurons [188]. As different from the frog sciatic nerve, topiramate inhibited TTX-sensitive Na+-channels with an IC50 value of 0.0489 mM in rat cerebellar granule cells [189]. Such a distinction would be possibly attributed to a distinction in topiramate sensitivity among different types or phosphorylation states of Na+ channels [190]. Antinociceptive actions of sodium valproate and topiramate have been attributed to other mechanisms including GABAA-receptor response increase (see [191][192] for review). Glutamate-receptor inhibition also would possibly contribute to the antinociceptions produced by topiramate and lamotrigine, because topiramate depresses GluK1 (GluR5) kainate receptors (a subtype of glutamate receptors) in rat basolateral amygdala neurons [193] and lamotrigine suppresses α-amino-3-hydroxy-5-methyl-4-isoxazole propionate (AMPA) receptors (another subtype of glutamate receptors) in rat dentate gyrus granule cells [194]. Table 1 summarizes the actions of antiepileptics on frog sciatic nerve fast-conducting CAPs together with their IC50 values (see also [68]).

Antiepileptics having an ability to suppress frog sciatic nerve CAPs appeared to have antinociceptive actions in a persistent pain model. Intraperitoneal application of lamotrigine, carbamazepine and oxcarbazepine produced analgesic effects in the second phase of the formalin test (that reflects inflammation occurring 15-20 min after formalin injection) whereas phenytoin, topiramate and sodium valproate did not so in rats [195][196].  The antinociceptive effects of antiepileptics appeared to be partly related to nerve AP conduction inhibition produced by them. The plasma concentrations of lamotrigine and carbamazepine used to clinically treat epilepsy are, respectively, < 12 μM and 20-50 μM [197][198], values smaller than those of IC50 for frog sciatic nerve CAP inhibition.

3.3. Antidepressants

Antidepressants are thought to alleviate pain by activating the 5-HT- and NA-containing descending antinociceptive pathway to the spinal dorsal horn from brainstem through 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]; see [205][206] for review) and the inhibition of voltage-gated Ca2+ [207][208], NMDA-receptor [209][210][211][212] and P2X4-receptor channels (a subtype of ionotropic P2X receptors; [213]), all of which are related to synaptic transmission. Moreover, an inhibition of neuroimmune mechanisms accompanying nerve injury may be involved in pain alleviation produced by antidepressants [214].

A 5-HT and NA reuptake inhibitor (SNRI) duloxetine ([215][216][217]; see [218] for review) inhibited frog sciatic nerve CAPs in a partially reversible manner. Duloxetine activity was concentration-dependent in a range of 0.001-2 mM with an IC50 value of 0.23 mM [33]. A similar CAP inhibition was produced by a selective 5-HT reuptake inhibitor (SSRI) fluoxetine ([201][202]; see [206][219] for review). Fluoxetine-induced CAP peak amplitude reduction was partially reversible, concentration-dependent in a range of 0.05-5 mM and had an IC50 of 1.5 mM, a value larger than that of duloxetine [33].

Typical tricyclic antidepressants (amitriptyline and desipramine, which are tertiary and secondary amines, respectively; [200][203][220][221]) also inhibited frog sciatic nerve CAPs. Amitriptyline reduced CAP peak amplitudes in a concentration range of 0.001-1 mM with an IC50 value of 0.26 mM, and desipramine did so in a concentration range of 0.1-5 mM with an IC50 value of 1.6 mM [33]. Thus, amitriptyline was six-fold more effective in CAP inhibition than desipramine. Consistent with this result, amitriptyline produced an AP conduction blockade in the rat sciatic nerve [156]. Like tricyclic antidepressants, a tetracyclic one maprotiline [220] also inhibited frog sciatic nerve CAPs in a partially reversible manner. Maprotiline activity was concentration-dependent in a range of 0.2-5 mM with an IC50 value of 0.95 mM [33]. Trazodone, 5-HT2-receptor antagonist and reuptake inhibitor (SARI), is a non-SNRI, -SSRI, -tricyclic and -tetracyclic antidepressant ([222][223][224][225]; see [226] for review). Trazodone reduced frog sciatic nerve CAP peak amplitudes at concentrations ranging from 0.2 to 2 mM, a maximally dissolved concentration, in a partially reversible manner. The extent of trazodone (1 mM)-induced CAP peak amplitude reduction was about 50% [33].

The antidepressants-induced CAP inhibition would be due to an attenuation of TTX-sensitive voltage-gated Na+ channels which are involved in the production of frog sciatic nerve CAPs. Consistent with this idea, voltage-gated Na+ channels were inhibited by duloxetine [227][228], fluoxetine [229], amitriptyline [227][229][230][231][232][233][234][235], desipramine and maprotiline [236]. TTX-sensitive voltage-gated Na+ channels in bovine adrenal chromaffin cells were suppressed by amitriptyline (IC50 = 0.0202 mM), fluoxetine (62% amplitude reduction at 0.02 mM), desiparmine (50% at 0.02 mM) and trazodone (20% at 0.1 mM; [229]). Amitriptyline also reduced Na+-channel current amplitudes in rat clonal pituitary GH3 cells with an IC50 value of 0.0398 mM [156]. These efficacies for Na+-channel inhibition were much larger than those of frog sciatic nerve CAPs. Furthermore, IC50 value (0.0221 mM) for duloxetine to inhibit (TTX-sensitive) Nav1.7 channels was about 10-fold less than that (0.23 mM) for frog sciatic nerve CAP inhibition [228], while the efficacy sequence for CAP inhibition (maprotiline > fluoxetine) was the same as that of the Nav1.7 channel, where IC50 values for maprotiline, fluoxetine, desipramine and amitriptyline were 0.028, 0.074, 0.024 and 0.085 mM, respectively [236]. The observation that amitriptyline and duloxetine had a similar IC50 value for frog sciatic nerve CAP inhibition was the same as that for cardiac-type Na+-channel inhibition [227]. TTX-resistant Na+ channel, possibly Nav1.8 channel, current amplitude in rat trigeminal ganglion neurons was also reduced by amitriptyrine with an IC50 value of 0.00682 mM [237]. With respect to chemical structure, typical local anesthetics have hydrophilic and hydrophobic moieties that are separated by an intermediate amide or ester linkage (see [238] for review), while all of the antidepressants tested in the frog sciatic nerve, except for trazodone, have a hydrophilic amine group and a hydrophobic moiety containing benzene rings, both of which are linked by a straight chain hydrocarbon (see Fig. 1 in

[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.

for the chemical structures of six antidepressants tested). Such chemical structures may take a pivotal role in inhibiting Na+ channels. Table 1 summarizes IC50 values for antidepressants to inhibit frog sciatic nerve fast-conducting CAPs (see also [68]).

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 example, duloxetine inhibited tactile allodynia (where pain is caused by a stimulus that does not normally elicit pain) and heat hyperalgesia (an abnormally increased sensitivity to pain) in neuropathic pain rat models [215]. Fluoxetine produced analgesia in streptozotocin-induced diabetic neuropathic pain mouse models [201]. Amitriptyline and desipramine were effective in attenuating pain in patients with diabetic neuropathy [200]. Maprotiline suppressed neuropathic pain produced by chronic constriction injury of the sciatic nerve in rats [241]. Trazodone depressed hyperalgesia in chronic constriction injury rat models [222]. The plasma concentrations of duloxetine, fluoxetine, amitriptyline, desipramine, maprotiline and trazodone used to clinically treat depression and neuropathic pain are, respectively, 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 [205][228]. These concentration values were much smaller than those of IC50 for frog sciatic nerve CAP inhibition. The antidepressants may produce an antinociception only when applied locally to the nerve.

3.4. Adrenoceptor Agonists

Epidural and intrathecal administration of α2 adrenoceptor agonists including clonidine and DEX (see [242] for review) results in antinociception in animals [243][244][245] and humans [246]. This is possibly owing to the inhibition produced by the agonists of glutamatergic excitatory transmission in spinal superficial dorsal horn neurons [247]. α2 Adrenoceptor agonists combined with local anesthetics in spinal anesthesia lead to the extension of peripheral nerve block duration in animals [248][249][250][251] and humans ([252][253][254][255][256][257][258][259]; see [260] for review). This is possibly due to a local vessel contraction produced by the agonists, leading to a decrease in the clearance of the anesthetics from the subarachnoid space [261][262]. Moreover, α2 adrenoceptor agonists attenuate nerve AP conduction and therefore have a local anesthetic effect, contributing to enhanced local anesthetic effect [263]. For instance, clonidine not only suppressed excitatory transmission in rat spinal lamina II neurons [264][265] but also blocked AP conduction in peripheral nerves [172][263][266]. The latter action required a much higher concentration of clonidine than the former one. DEX as well as clonidine reportedly inhibited excitatory transmission in rat lamina II neurons [267]. Intracutaneous application of DEX or clonidine together with lidocaine into the back of guinea-pigs increased the extent of the local anesthetic effect of lidocaine [268]. Local wound infiltration with DEX added to another local anesthetic bupivacaine more effectively attenuated postoperative pain compared to bupivacaine alone in humans [269]. In addition to clonidine, DEX possibly has an inhibitory action on nerve AP conduction, because DEX is reported to suppress voltage-gated Na+-channels [146] (see below).

Frog sciatic nerve CAP peak amplitudes were reduced by DEX in a concentration-dependent manner in a range of 0.01-1 mM with an IC50 value of 0.40 mM [34]. The DEX activity was not inhibited by α2-adrenoceptor antagonists, yohimbine and atipamezole ([247][270][271][272]; see [273][274] for review), although DEX exhibited a high affinity for α2 adrenoceptors [242]. This result indicates no involvement of α2 adrenoceptors in the DEX activity [34]. CAP peak amplitude reduction was also seen by other α2-adrenoceptor agonist, oxymetazoline, which is more selective to α2A than α2B and α2C (see [273][275] for review), and also by clonidine in a manner resistant to yohimbine. Oxymetazoline reduced CAP peak amplitude with an IC50 value of 1.5 mM. Clonidine at 2 mM reduced CAP peak amplitude by about 20% [34]. This clonidine activity was different in extent from that (CAP amplitude reduction of 80% at 0.3 mM) reported by Starke et al. [172] for frog sciatic nerve CAPs, although a reason for this discrepancy was unknown. On the other hand, frog sciatic nerve CAPs were not affected by various adrenoceptor agonists, adrenaline, NA, α1-adrenoceptor agonist phenylephrine and β-adrenoceptor agonist isoproterenol at 1 mM [34]. A similar clonidine-induced CAP inhibition has been reported in the rat sciatic nerve. Thus, CAPs originating from primary-afferent Aα and C fibers in the rat sciatic nerve were found to be suppressed by clonidine with IC50 values of 2.0 and 0.45 mM, respectively [266].

The α2-adrenoceptor agonists-induced CAP inhibition would be mediated by a suppression of voltage-gated Na+ and K+ channels involved in AP production. DEX reportedly reduced voltage-gated Na+ channel current amplitude in rat DRG neurons in a manner insensitive to yohimbine, although this type of Na+ channels was resistant to TTX [146]. IC50 value (0.058 mM) for this rat DRG neuron DEX activity was about 10-fold smaller than that (0.40 mM) for frog sciatic nerve CAP inhibition. The rat TTX-resistant Na+ channel current amplitude was also reduced by clonidine with an IC50 value of 0.26 mM [146]. TTX-sensitive Na+ channel current amplitudes in DRG neuroblastoma hybridoma cell line ND7/23 cells were reduced by clonidine (IC50 = 0.824 mM; [155]). It has been reported in NG108-15 neuronal cells that delayed-rectifier K+-channel current amplitudes are reduced by DEX with an IC50 value of 0.0046 mM while TTX-sensitive Na+-channel current amplitudes are reduced by about 20% by DEX (0.01 mM) in a manner insensitive to yohimbine [276]. These differences in drug potency may be due to a difference in either Na+-channel types or animal species. Table 1 summarizes the actions of adrenoceptor agonists on frog sciatic nerve fast-conducting CAPs together with their IC50 values (see also [68]).

In clinical practice, DEX administration results in producing analgesia/sedation and decreasing heart rate, cardiac output and memory, each of whose effects depends on the plasma concentration of DEX in a different manner [277]. In patients, sedation is rapidly induced by 0.2 to 0.7 mg·kg-1·hr-1 i.v. [242]. In intramuscular application in cats, 40 mg·kg-1 is a usual dose to produce analgesia/sedation [278]. DEX concentrations necessary to inhibit frog sciatic nerve AP conduction are >1000-fold higher than those of the usage of DEX as α2 adrenoceptor agonist, because the clinical usage of DEX is < 0.05 μM for plasma levels (see [277]). Therefore, the DEX’s potency in blocking nerve AP conduction is independent of the usage of DEX for analgesia/sedation. α2-Adrenoceptor agonists including DEX, combined with a local anesthetic, have been used to extend peripheral nerve AP conduction block duration [249][252][253][254][255][279]. It is possible that this effect is mediated by a local vasoconstriction leading to a delay of the absorption of the local anesthetic and/or a direct nerve AP conduction suppression produced by α2 adrenoceptor agonists [256]. The latter mechanism would be the above-mentioned nerve CAP inhibition produced by α2 adrenoceptor agonists. This action makes sense when considering their topical administration on nerves, but is not related to their usage for analgesia/anesthesia by systemic application. A chemical structure related to their α2 adrenoceptor agonists (see [34]) may play a crucial role in producing nerve AP conduction blockage.

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

As noted from Table 1, some of analgesic adjuvants had similar IC50 values for frog sciatic nerve CAP inhibitions. For example, antidepressants had IC50 values similar to those of some of local anesthetics, antiepileptics and α2-adrenoceptor agonists. Duloxetine and amitriptyline values (0.23 and 0.26 mM, respectively) were similar to those of ropivacaine, levobupivacaine, pramoxine, lamotrigine, carbamazepine and DEX (0.34, 0.23, 0.21, 0.44, 0.50 and 0.40 mM, respectively). On the other hand, fluoxetine, desipramine, maprotiline and trazodone values (1.5, 1.6, 0.95 and ca. 1 mM, respectively) were similar 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: 0.2-0.5 mM) or latter (IC50: 0.7-2 mM) drugs, although the number of CH2 in opioids having similar structures was related to the extent of CAP inhibition (see Section 2.2). The antidepressants had IC50 values being much larger than that of tetracaine (0.014 mM). Thus, some of the analgesic adjuvants will have an ability to inhibit nerve AP conduction with an efficacy comparable to each other.

When IC50 values of analgesic adjuvants were compared with those of antipyretic analgesics NSAIDs, diclofenac’s IC50 value (0.94 mM) was close to those of lidocaine, cocaine, maprotiline and trazodon, (0.74, 0.80, 0.95 and ca. 1 mM, respectively), while aceclofenac, tolfenamic acid, meclofenamic acid and flufenamic acid (0.47, 0.29, 0.19 and 0.22 mM, respectively) had IC50 values being similar to those of ropivacaine, levobupivacaine, pramoxine, duloxetine, amitriptyline, lamotrigine, carbamazepine and DEX (0.34, 0.23, 0.21, 0.23, 0.26, 0.44, 0.50 and 0.40 mM, respectively). The NSAIDs’ values were smaller than those of procaine, prilocaine, fluoxetine, desipramine and oxymetazoline (2.2, 1.8, 1.5, 1.6 and 1.5 mM, respectively), while being larger than that of tetracaine (0.014 mM). Thus, NSAIDs could suppress nerve AP conduction with efficacies comparable to some of analgesic adjuvants. In these cases, there were no common chemical structures among compounds having similar IC50 values.

Not only NSAIDs and analgesic adjuvants but also narcotic analgesics opioids have an ability to inhibit nerve AP conduction. Frog sciatic nerve CAP peak amplitudes were also reduced by opioids; tramadol and ethylmorphine had the IC50 values of 2.3 and 4.6 mM, respectively; morphine and codeine at 5 mM reduced CAP amplitude by 15% and 30%, respectively. These opioid actions were smaller in extent than those of NSAIDs and analgesic adjuvants. For instance, the IC50 value (2.3 mM) of tramadol was larger by 3.1- and 6.8-fold than those (0.74 mM and 0.34 mM, respectively) of lidocaine and ropivacaine, respectively [27]. Lidocaine is previously reported by other investigators to reduce frog sciatic nerve CAP amplitudes with an IC50 of 6.6 mM, a value larger by three-fold than the tramadol value [107]. Ratio of the IC50 value of tramadol to that of lidocaine was almost comparable to Katsuki et al. [27]’s one, albeit IC50 values for lidocaine were largely different between the two studies [27][107]. A contribution of nerve AP conduction inhibition to analgesia produced by opioids appears to be much less in extent compared to well-known other cellular mechanisms including a membrane hyperpolarization and a decrease in the release of L-glutamate from nerve terminals in the spinal dorsal horn (for instance, see [74][75]). In conclusion, nerve AP conduction suppression may be a common mechanism for antinociception produced by NSAIDs and analgesic adjuvants but not opioids.

IC50 values for the analgesic adjuvants to inhibit frog sciatic nerve CAPs were similar to those in rat sciatic nerve CAPs while being generally larger than IC50 values for TTX-sensitive Na+ channel inhibitions. This difference could be explained by several possibilities. First, CAPs are produced by not only voltage-gated Na+ but also K+ channels. Second, TTX-sensitive Na+-channel types (Nav1.1-1.4, Nav1.6 and Nav1.7) may differ in expression among the preparations examined. Third, CAPs originate from a bundle of nerve fibers while Na+ currents from single cells. When the analgesic adjuvants used clinically act on the nerve trunk, their sciatic nerve IC50 values may be an appropriate measure for a nerve AP conduction inhibition in vivo, because nerve conduction is mediated by both voltage-gated Na+ and K+ channels. Considering that nociceptor-specific deletion of TTX-sensitive Nav1.7 gene results in inhibited acute and inflammatory pain in mice [280], Na+ channels may be the main target of analgesics and analgesic adjuvants.

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

Many of plant-derived compounds activate TRP channels located in the peripheral terminals of primary-afferent Aδ-fiber and C-fiber neurons, resulting in the production of AP, which in turn leads to temperature sensation and nociception. For example, capsaicin, allyl isothiocyanate and menthol activate TRPV1, TRPA1 and TRP melastatin-8 (TRPM8) channels, respectively (for example, see [15][281][282] for review). On the other hand, TRPV1, TRPA1 and TRPM8 channels are expressed in primary-afferent neuron central terminals in lamina II of the spinal dorsal horn, and 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 ([283][284][285][286][287][288][289][290]; see [291][292] for review). This activation is thought to be involved in the modulation of excitatory and inhibitory synaptic transmission in lamina II neurons, resulting in nociceptive transmission modulation. This idea is supported by the observation that synaptic transmission in lamina II neurons is affected by a variety of endogenous pain modulators (for example, see [17] for review).

As with analgesics and analgesic adjuvants, CAPs in the frog sciatic nerve were inhibited by many of plant-derived compounds that produce antinociception by their topical, oral, intraperitoneal or intrathecal administration (see [293][294] for review). 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 owing to its less solubility, capsaicin at 0.1 mM reduced CAP peak amplitudes by 36% (Table 1); this action could be attributed to at least a part of alleviation of chronic pain produced by capsaicin applied to the skin ([295][296]; for example, see [15] for review). Their plant-derived compounds’ activities were close to those of analgesic adjuvants and NSAIDs ([32][297][298][299]; see [35] for review). Thus, their IC50 values were comparable to those of duloxetine (0.23 mM), amitriptyline (0.26 mM), aceclofenac (0.47 mM), tolfenamic acid (0.29 mM), meclofenamic acid (0.19 mM) and flufenamic acid (0.22 mM). 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 the IC50 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); these values were similar to those of fluoxetine (1.5 mM) and desipramine (1.6 mM). Linalyl acetate, eugenol, (-)-menthol and (+)-menthol had the 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 (0.94, 0.95, and ca. 1.0 mM, respectively).

The cinnamaldehyde and allyl isothiocyanate activities were resistant to a non-selective TRP antagonist ruthenium red, indicating no involvement of TRP channels [298]. Capsaicin at high concentrations (0.03-0.1 mM) inhibited voltage-gated Na+-channels in rodents in a manner independent of TRPV1 channels [300][301][302]. Other plant-derived compounds-induced CAP inhibitions are possibly mediated by an attenuation of TTX-sensitive voltage-gated Na+ channels (for example, eugenol [303], thymol [304], carvacrol [305] and linalool [306]; see [307] for review). IC50 value (0.37 mM) for inhibition by carvacrol of voltage-gated Na+ channels in rat DRG neurons [308] 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) had no effect on CAPs (Table 1). Their compounds’ activities were much smaller than those of NSAIDs and analgesic adjuvants. In conclusion, some plant-derived compounds could be used instead of NSAIDs and analgesic adjuvants in terms of nerve AP conduction inhibition. Plant-derived compounds will be 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 compound hinokitiol (β-thujaplicin; 2-hydroxy-4-isopropylcyclohepta-2,4,6-trien-1-one; contained in a species of cypress tree [309]) also reduced frog sciatic nerve CAP peak amplitudes with an IC50 value of 0.54 mM (Table 1). This value was comparable to those of many other plant-derived compounds. The CAP inhibition is possibly mediated by an interaction involving the carbonyl, isopropyl and hydroxyl groups of hinokitiol [310]. The carbonyl group serves for the seven-membered ring of hinokitiol to act as a benzene ring, while the isopropyl and hydroxyl groups are important for the CAP amplitude reduction produced by hinokitiol. This idea is supported by the observation that benzene-ring compounds having the isopropyl and hydroxyl groups, such as thymol, carvacrol, biosol (the last of which is a stereoisomer of thymol and carvacrol; IC50 = 0.58 mM), and also 4-isopropylphenol (0.85 mM) had an ability to suppress frog sciatic nerve CAPs (see above; [297][310]). As with other plant-derived compounds, hinokitiol has a variety of actions including inhibition of apoptosis [311], anti-bacterial, anti-inflammatory [312], insecticidal [313], anti-fungal [314], anti-tumor [315] and cytotoxic activities [316][317]. Hinokitiol used as a dermatological drug to inhibit inflammation may have a local anesthetic effect. Application of an oral care gel containing hinokitiol to the oral mucosa is reported to alleviate oral pain in patients with oral lichen planus related to hepatitis C virus infection [318]. Such a pain attenuation may be partly mediated by a local anesthetic effect of hinokitiol

A CAP inhibitory action similar to hinokitiol was seen with a general anesthetic propofol (2,6-diisopropylphenol; [319][320][321]; see [322][323] for review) having two isopropyl groups and one hydroxyl group bound to the benzene ring. Thus, propofol reduced frog sciatic nerve CAP peak amplitudes in a concentration-dependent manner with an IC50 of 0.14 mM, a value smaller than that of hinokitiol (Table 1; [30]). Consistent with this observation, propofol reportedly suppressed APs recorded extracellularly in the human and mammalian CNS [324][325]. 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 lamina II neurons [326].

Traditional Japanese (Kampo) medicines composed of plant-derived crude chemicals are used together with Western medicines in Japan with various purposes such as antinociception (see [327][328][329][330][331][332] for review). 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 reduced by daikenchuto and also by other Kampo medicines, rikkosan, kikyoto, rikkunshito, shakuyakukanzoto and kakkonto in a concentration-dependent manner. Among these medicines, daikenchuto was the most effective with an IC50 value of 1.1 mg/mL. Daikenchuto has three kinds of crude medicine, Japanese pepper, processed ginger and ginseng radix, the former two of which suppressed CAPs while the last had no effects on CAPs. Japanese pepper had an IC50 value of 0.77 mg/mL and processed ginger at 2 mg/mL reduced CAP peak amplitudes by 31% [333]. A small part of the analgesic effect of Kampo medicine may be due to its nerve AP conduction inhibitory action.

Table 1. Effects of analgesics, analgesic adjuvants and plant-derived compounds on fast-conducting CAPs recorded from the frog sciatic nerve.

(1) NSAIDs [26]

(a) Acetic acid-based NSAIDs

Diclofenac

Aceclofenac

Indomethacin

Acemetacin

Etodolac

   

0.94 mM

0.47 mM

38% reduct. (1 mM) 

38% reduct. (1 mM) 

15% reduct. (1 mM)

   

Sulindac

Felbinac

      

No effect (1 mM)

No effect (1 mM)

      

(b) Fenamic acid-based NSAIDs

Tolfenamic acid

Meclofenamic acid

Mefenamic acid

Flufenamic acid

    

0.29 mM

0.19 mM

16% reduct. (0.2 mM)

0.22 mM

    

(c) Salicylic acid-based NSAID

Aspirin

       

No effect (1 mM)

       

(d) Propionic acid-based NSAID

Ketoprofen

Naproxen

Ibuprofen

Loxoprofen

Flurbiprofen

   

No effect (1 mM)

No effect (1 mM) 

No effect (1 mM)

No effect (1 mM)

No effect (1 mM)

   

(e) Enolic acid-based NSAID

Meloxican

Piroxicam

      

No effect (0.5 mM) 

No effect (1 mM)

      

(2) Opioids [27][28]

Tramadol

Mono-O-desmethyl-tramadol

Morphine

Codeine

Ethylmorphine

   

2.3 mM

9% reduct. (5 mM)

15% reduct. (5 mM)

30% reduct. (5 mM)

4.6 mM

   

(3) Local anesthetics [27][28][29][30][31][32]

(a) Amide-type local anesthetics

Lidocaine

Ropivacaine

Prilocaine

Levobupivacaine

Bupovacaine

   

0.74 mM

0.34 mM

1.8 mM

0.23 mM

76% reduct. (0.5 mM)

   

(b) Ester-type local anesthetics

Cocaine

Procaine

Benzocaine

Tetracaine

    

0.80 mM

2.2 mM

0.80 mM

0.014 mM

    

(c) Other-type local anesthetic

Pramoxine

       

0.21 mM

       

(4) Antiepileptics [29]

Lamotrigine

Carbamazepine

Oxcarbazepine

   Phenytoin

Gabapentin

Topiramate

Sodium valproate

 

0.44 mM

0.50 mM

20% reduct. 

15% reduct.

No effect 

No effect

No effect

 
  

(0.5 mM)

(0.1 mM)

(10 mM)

(10 mM)

(10 mM)

 

(5) Antidepressants [33]

Duloxetine

Fluoxetine

Amitriptyline

Desipramine

Maprotiline

Trazodone

  

0.23 mM

1.5 mM

0.26 mM

1.6 mM

0.95 mM

ca. 1.0 mM

  

(6) Adrenoceptor agonists [34]

Adrenaline

Noradrenaline

Dexmedetomidine

Oxymetazoline

Clonidine

Phenylephrine

Isoproterenol

 

No effect

No effect

0.40 mM

1.5 mM 

ca. 20% reduct.

No effect 

No effect

 

(1 mM)

(1 mM) 

  

(2 mM)

(1 mM)

(1 mM)

 

(7) Plant-derived compounds

(a) Open chain or six-membered ring compounds [32][297][298][299]

Carvacrol

Thymol

Citronellol

Bornyl acetate

Citral

Citronellal

Geranyl acetate

 

0.34 mM

0.34 mM

0.35 mM

0.44 mM

0.46 mM

0.50 mM

0.51 mM

 

Geraniol

Capsaicin    

(+)-Pulegone

(-)-Carvone

(+)-Carvone  

(+)-Borneol

 (±)-Linalool

 

0.53 mM

36% reduct. (0.1 mM) 

1.4 mM

1.4 mM

2.0 mM

1.5 mM

    1.7 mM

 

(-)-Menthone

(+)-Menthone 

(-)-Carveol  

α-Terpineol 

Rose oxide 

Cinnamaldehyde 

Allyl isothiocyanate

 

1.5 mM  

2.2 mM 

1.3 mM

2.7 mM 

2.6 mM 

1.2 mM 

1.5 mM 

 

Linalyl acetate 

Eugenol 

(-)-Menthol

(+)-Menthol 

1,8-Cineole

1,4-Cineole 

Zingerone

Guaiacol

0.71 mM

0.81 mM 

1.1 mM

0.93 mM 

    5.7 mM

7.2 mM

8.3 mM

7.7 mM 

Vanillin

p-Cymene

  Myrcene 

Vanillylamine

  (+)-Limonene

   

9.0 mM

22% reduct. (2 mM)

7% reduct. (5 mM)

   12% reduct. (10 mM)

8% reduct. (10 mM)

   

Vanillic acid

p-Menthane 

Menthyl chloride

     

No effect (7 mM)

No effect (10 mM)

No effect (10 mM)

     

(b) Seven-membered ring compound [310]

Hinokitiol

       

0.54 mM

       

(8) General anesthetic [30]

Propofol

       

0.14 mM

       

Here, IC50 value for CAP inhibition and the extent of CAP amplitude reduction (reduct.) at each concentration are shown.

6. Conclusion

This research demonstrated that frog sciatic nerve CAPs are depressed by some of NSAIDs, analgesic adjuvants and plant-derived compounds having analgesic activities with similar efficacies and also by opioids with efficacies less than them. Although the CAPs are originated from fast-conducting TTX-sensitive Aα fibers, nociceptive information is transferred through slow-conducting Aδ and C fibers [1]. Slow-conducting CAPs were not able to be recorded from the frog sciatic nerve, because Aδ-fiber CAPs were not able to be isolated from Aα-fiber ones and C-fiber CAPs were much smaller in peak amplitude and conduction velocity than fast-conducting Aα ones [25]. Therefore, the effects of the antinociceptive compounds on slow-conducting CAPs were not able to be examined. In order to know a detail of differences in nerve AP conduction inhibition extent among various antinociceptive compounds, it would be necessary to examine their effects on slow-conducting CAPs.

In nerves other than the frog sciatic nerve, antinociceptive drugs reportedly suppress not only A-fiber but also C-fiber CAPs. For instance, in the rabbit vagus nerve, fentanyl and sufentanil inhibited C-fiber CAPs with extents smaller than those of A-fiber ones [86] and lidocaine blocked nerve AP conduction in not only myelinated A-fibers but also unmyelinated C-fibers ([334]; lidocaine inhibited more effectively rat A-fiber CAPs than C-fiber ones [335]). Clonidine inhibited both Aα-fiber and C-fiber CAPs (see Section 3.4). Furthermore, many researchers have demonstrated a suppression by NSAIDs [58][65][336], lidocaine and α2-adrenoceptor agonists [146] of TTX-resistant voltage-gated Na+ channels that may be involved in slow-conducting CAP production. Knockdown of TTX-resistant Nav1.8 channels reportedly results in inhibition of neuropathic and inflammatory pain in rats [337].

Since analgesic and analgesic adjuvant concentrations needed for CAP inhibition are higher than clinically relevant ones, as mentioned above, nerve AP conduction inhibition may occur only when the drugs are applied locally or are accumulated in the nervous system. If such a suppression occurs in Aα fibers innervating skeletal muscle, this would result in undesirable side effects such as muscle paralysis. Thus, the drugs will have to be used at the lowest possible concentrations. Primary-afferent C and Aδ fibers are smaller in diameter than Aα ones. Therefore, if the drugs act on voltage-gated Na+ channels from cytoplasm side, C fibers will precede Aα fibers in attenuating nerve AP conduction due to a distinction in surface-to-volume ratio between the fibers. At least a part of antinociception produced by analgesics and analgesic adjuvants could be attributed to their inhibitory actions on nerve AP conduction mediated by the activation of voltage-gated TTX-sensitive and TTX-resistant voltage-gated Na+ channels.

References

  1. Fields, H.L. Pain; McGraw-Hill: New York, NY, USA, 1987. Fields, H.L. Pain; McGraw-Hill: New York, U.S.A., 1987.
  2. Willis, W.D., Jr.; Coggeshall, R.E. Sensory Mechanisms of the Spinal Cord, 2nd ed.; Plenum: New York, NY, USA, 1991. Willis, W.D. Jr.; Coggeshall, R.E. Sensory Mechanisms of the Spinal Cord, 2nd ed.; Plenum: New York, U.S.A., 1991.
  3. Todd, A.J. Neuronal circuitry for pain processing in the dorsal horn. Nat. Rev. Neurosci. 2010, 11, 823–836. Todd, A.J. Neuronal circuitry for pain processing in the dorsal horn. Nat. Rev. Neurosci. 2010, 11, 823-836.
  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. 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.
  5. Merskey, H. Clarifying definition of neuropathic pain. Pain 2002, 96, 408–409. Merskey, H. Clarifying definition of neuropathic pain. Pain 2002, 96, 408-409.
  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. 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.
  7. Finnerup, N.B.; Sindrup, S.H.; Jensen, T.S. The evidence for pharmacological treatment of neuropathic pain. Pain 2010, 150, 573–581. Finnerup, N.B.; Sindrup, S.H.; Jensen, T.S. The evidence for pharmacological treatment of neuropathic pain. Pain 2010, 150, 573-581.
  8. Jensen, T.S. Anticonvulsants in neuropathic pain: Rationale and clinical evidence. Eur. J. Pain 2002, 6, 61–68. Jensen, T.S. Anticonvulsants in neuropathic pain: rationale and clinical evidence. Eur. J. Pain 2002, 6 (Suppl. A), 61-68.
  9. Kamibayashi, T.; Maze, M. Clinical uses of α2-adrenergic agonists. Anesthesiology 2000, 93, 1345–1349. Kamibayashi, T., Maze, M. Clinical uses of α2-adrenergic agonists. Anesthesiology 2000, 93, 1345-1349.
  10. Lynch, M.E. Antidepressants as analgesics: A review of randomized controlled trials. J. Psychiatry Neurosci. 2001, 26, 30–36. Lynch, M.E. Antidepressants as analgesics: a review of randomized controlled trials. J. Psychiatry Neurosci. 2001, 26, 30-36.
  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. 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.
  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. 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.
  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. 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.
  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. 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.
  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. 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.
  16. Fürst, S. Transmitters involved in antinociception in the spinal cord. Brain Res. Bull. 1999, 48, 129–141. Fürst, S. Transmitters involved in antinociception in the spinal cord. Brain Res. Bull. 1999, 48, 129-141.
  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. 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.
  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. Zeilhofer, H.U.; Wildner, H.; Yévenes, G.E. Fast synaptic inhibition in spinal sensory processing and pain control. Physiol. Rev. 2012, 92, 193-235.
  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. 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.
  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. 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.
  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. 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.
  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. Kiernan, M.C.; Bostock, H.; Park, S.B.; Kaji, R.; Krarup, C.; Krishnan, A.V.; Kuwabara, S.; Lin, C.S.; Misawa, S.; Moldovan, M.; Sung, J.; Vucic, S.; Wainger, B.J.; Waxman, S.; Burke, D. Measurement of axonal excitability: consensus guidelines. Clin. Neurophysiol. 2020, 131, 308-323.
  23. Levitan, I.B.; Karczmarek, L.K. The Neuron, 3rd ed.; Oxford University Press: New York, NY, USA, 2002. Levitan, I.B.; Karczmarek, L.K. The Neuron, 3rd ed.; Oxford University Press: New York, U.S.A., 2002.
  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. 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, U.S.A., 2012, pp. 89-106.
  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. Kobayashi, J.; Ohta, M.; Terada, Y. C fiber generates a slow Na+ spike in the frog sciatic nerve. Neurosci. Lett. 1993, 162, 93-96.
  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. 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.
  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. 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.
  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. 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.
  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. 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.
  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. 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.
  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. 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.
  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. 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.
  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. 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.
  34. 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.
  35. Kumamoto, E. Effects of plant-derived compounds on excitatory synaptic transmission and nerve conduction in the nervous system – involvement in pain modulation. Curr. Top. Phytochemistry 2018, 14, 45-70.
  36. Ferreira, S.H. Prostaglandins, aspirin-like drugs and analgesia. Nat. New Biol. 1972, 240, 200-203.
  37. 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.
  38. Vane, J.R. Introduction: mechanism of action of NSAIDs. Br. J. Rheumatol. 1996, 35 (Suppl. 1), 1-3.
  39. Simmons, D.L.; Botting, R.M.; Hla, T. Cyclooxygenase isozymes: the biology of prostaglandin synthesis and inhibition. Pharmacol. Rev. 2004, 56, 387-437.
  40. 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, U.S.A., 2011, pp. 959-1004.
  41. 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.
  42. Garg, P.; Sanguinetti, M.C. Structure-activity relationship of fenamates as Slo2.1 channel activators. Mol. Pharmacol. 2012, 82, 795-802.
  43. 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.
  44. 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.
  45. 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.
  46. 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.
  47. 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.
  48. Guinamard, R.; Simard, C.; Del Negro, C. Flufenamic acid as an ion channel modulator. Pharmacol. Ther. 2013, 138, 272-284.
  49. 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.
  50. 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.
  51. 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.
  52. 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.
  53. Silva, L.C.R.; Castor, M.G.Me., 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.
  54. 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.
  55. Silva, L.C.R.; Castor, M.G.Me.; 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.
  56. Fowler, C.J. NSAIDs: eNdocannabinoid stimulating anti-inflammatory drugs? Trends Pharmacol. Sci. 2012, 33, 468-473.
  57. 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.
  58. 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.
  59. 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. Invest. Ophthalmol. Vis. Sci. 2007, 48, 182-188.
  60. 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.
  61. 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.; Han, J.; Park, J.-Y.; Hong, S.-G. 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.
  62. 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.
  63. 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.
  64. 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.
  65. 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.
  66. 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.
  67. Chen, X.; Gallar, J.; Belmonte, C. Reduction by antiinflammatory drugs of the response of corneal sensory nerve fibers to chemical irritation. Invest. Ophthalmol. Vis. Sci. 1997, 38, 1944-1953.
  68. Kumamoto, E. Inhibition of fast nerve conduction produced by analgesics and analgesic adjuvants – possible involvement in pain alleviation. Pharmaceuticals 2020, 13, 62.
  69. 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.
  70. 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.
  71. 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. Pflügers Arch. 2010, 459, 579-592.
  72. 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.
  73. 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.
  74. 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.
  75. 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.
  76. Yoshimura, M.; North, R.A. Substantia gelatinosa neurones hyperpolarized in vitro by enkephalin. Nature 1983, 305, 529-530.
  77. North, R.A. Opioid actions on membrane ion channels. In: Handbook of Experimental Pharmacology, Vol. 104; Herz, A., Ed.; Springer: Berlin, Germany, 1993, pp. 773-797.
  78. Yaksh, T.L. Pharmacology and mechanisms of opioid analgesic activity. Acta Anaesthesiol. Scand. 1997, 41, 94-111.
  79. Smith, T.W.; Buchan, P.; Parsons, D.N.; Wilkinson, S. Peripheral antinociceptive effects of N-methyl morphine. Life Sci. 1982, 31, 1205-1208.
  80. 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.
  81. 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.
  82. Wenk, H.N.; Brederson, J.-D.; Honda, C.N. Morphine directly inhibits nociceptors in inflamed skin. J. Neurophysiol. 2006, 95, 2083-2097.
  83. 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.
  84. Stein, C.; Schäfer, M.; Machelska, H. Attacking pain at its source: new perspectives on opioids. Nature Med. 2003, 9, 1003-1008.
  85. 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.
  86. 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.
  87. 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.
  88. Jurna, I.; Grossmann, W. The effect of morphine on mammalian nerve fibres. Eur. J. Pharmacol. 1977, 44, 339-348.
  89. Coggeshall, R.E.; Zhou, S.; Carlton, S.M. Opioid receptors on peripheral sensory axons. Brain Res. 1997, 764, 126-132.
  90. 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.
  91. Wenk, H.N.; Honda, C.N. Immunohistochemical localization of delta opioid receptors in peripheral tissues. J. Comp. Neurol. 1999, 408, 567-579.
  92. Klotz, U. Tramadol - the impact of its pharmacokinetic and pharmacodynamic properties on the clinical management of pain. Arzneimittelforschung 2003, 53, 681-687.
  93. Lintz, W.; Erlacin, S.; Frankus, E.; Uragg, H. Metabolismus von Tramadol bei Mensch und Tier. Arzneimittelforschung 1981, 31, 1932-1943.
  94. Hennies, H.-H.; Friderichs, E.; Schneider, J. Receptor binding, analgesic and antitussive potency of tramadol and other selected opioids. Arzneimittelforschung 1988, 38, 877-880.
  95. 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.
  96. 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.
  97. 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.
  98. 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.
  99. 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.
  100. 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.
  101. 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.
  102. Le Roux, P.J.; Coetzee, J.F. Tramadol today. Curr. Opin. Anaesth. 2000, 13, 457-461.
  103. 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.
  104. 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.
  105. 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.
  106. 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.
  107. 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.
  108. 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.
  109. 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.
  110. 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.
  111. 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.
  112. 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.
  113. 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.
  114. 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.
  115. 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.
  116. Hunter, E.G.; Frank, G.B. An opiate receptor on frog sciatic nerve axons. Can. J. Physiol. Pharmacol. 1979, 57, 1171-1174.
  117. Kumamoto, E.; Mizuta, K.; Fujita, T. Opioid actions in primary-afferent fibers – involvement in analgesia and anesthesia. Pharmaceuticals 2011, 4, 343-365.
  118. 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.
  119. 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.
  120. 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.
  121. 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.
  122. 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.
  123. 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.
  124. Scholz, A. Mechanisms of (local) anaesthetics on voltage-gated sodium and other ion channels. Br. J. Anaesth. 2002, 89, 52-61.
  125. Hu, S.; Rubly, N. Effects of morphine on ionic currents in frog node of Ranvier. Eur. J. Pharmacol. 1983, 95, 185-192.
  126. 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.
  127. 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.
  128. 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, U.S.A., 2006, pp. 547-590.
  129. 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.
  130. 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. Nature Neurosci. 2001, 4, 268-274.
  131. Mays, K.S.; Lipman, J.J.; Schnapp, M. Local analgesia without anesthesia using peripheral perineural morphine injections. Anesth. Anal. 1987, 66, 417-420.
  132. 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. Therap. 1994, 271, 1528-1534.
  133. 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.
  134. 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.
  135. Hille, B. Ionic Channels of Excitable Membranes; Sinauer Associates Inc.: Massachusetts, 1984.
  136. 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.
  137. 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.
  138. 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.
  139. 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.
  140. 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.
  141. 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.
  142. 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.
  143. 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.
  144. 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. Invest. 2008, 118, 763-776.
  145. Leffler, A.; Lattrell, A.; Kronewald, S.; Niedermirtl, F.; Nau, C. Activation of TRPA1 by membrane permeable local anesthetics. Mol. Pain 2011, 7, 62.
  146. 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.
  147. 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. Devel. Ther. 2022, 16, 1171–1181.
  148. 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.
  149. McClellan, K.J.; Faulds, D. Ropivacaine: an update of its use in regional anaesthesia. Drugs 2000, 60, 1065-1093.
  150. 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.
  151. 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.
  152. 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.
  153. Foster, R.H.; Markham, A. Levobupivacaine: a review of its pharmacology and use as a local anaesthetic. Drugs 2000, 59, 551-579.
  154. 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.
  155. 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.
  156. Gerner, P.; Mujtaba, M.; Sinnott, C.J.; Wang, G.K. Amitriptyline versus bupivacaine in rat sciatic nerve blockade. Anesthesiology 2001, 94, 661-667.
  157. 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.
  158. 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.
  159. Carney, T.P. Alkaloids as local anesthetics. In: The Alkaloids, Vol. 5; Manske, R.H.F., Ed.; Academic Press: New York, U.S.A. 1955, pp. 211-227.
  160. Matthews, J.C.; Collins, A. Interactions of cocaine and cocaine congeners with sodium channels. Biochem. Pharmacol. 1983, 32, 455-460.
  161. 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.
  162. 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. Therap. 1996, 276, 568-577.
  163. 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.
  164. 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.
  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.
  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.
  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.
  168. Lee, H.-S. Recent advances in topical anesthesia. J. Dent. Anesth. Pain Med. 2016, 16, 237-244.
  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.
  170. Guénette, S.A.; Giroux, M.-C.; Vachon, P. Pain perception and anaesthesia in research frogs. Exp. Anim. 2013, 62, 87-92.
  171. Vanable, J.W. Benzocaine: an excellent amphibian anesthetic. Axolotl Newsletter 1985, 14, 19-21.
  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.
  173. Gissen, A.J.; Covino, B.G.; Gregus, J. Differential sensitivities of mammalian nerve fibers to local anesthetic agents. Anesthesiology 1980, 53, 467-474.
  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, U.S.A., 1997, pp. 1383-1391.
  175. Kubacka, M.; Rapacz, A.; Sałat, K.; Filipek, B.; Cios, A.; Pociecha, K.; Wyska, E.; Hubicka, U.; Żuromska-Witek, B.; Kwiecień, A.; Marona, H.; Waszkielewicz, A.M. 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.
  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. Pflügers Arch. 2001, 441, 425-433.
  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.
  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.
  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.
  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.
  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.
  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.
  183. Kuo, C.-C. A common anticonvulsant binding site for phenytoin, carbamazepine, and lamotrigine in neuronal Na+ channels. Mol. Pharmacol. 1998, 54, 712-721.
  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.
  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.
  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.
  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.
  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.; Pan, H.-L. The α2δ-1-NMDA Receptor complex Is critically involved in neuropathic pain development and gabapentin therapeutic actions. Cell Rep. 2018, 22, 2307-2321.
  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.
  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.
  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.
  192. Perucca, E. A pharmacological and clinical review on topiramate, a new antiepileptic drug. Pharmacol. Res. 1997, 35, 241-256.
  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.
  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.
  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.
  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.
  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.
  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, U.S.A., 1995, pp. 515-528.
  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.
  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.
  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.
  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.
  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.
  204. Hall, H.; Ögren, S.-O. Effects of antidepressant drugs on different receptors in the brain. Eur. J. Pharmacol. 1981, 70, 393-407.
  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, U.S.A., 2011, pp. 397-415.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  214. Kremer, M.; Yalcin, I.; Goumon, Y.; Wurtz, X.; Nexon, L.; Daniel, D.; Megat, S.; Ceredig, R.A.; Ernst, C.; Turecki, G.; Chavant, V.; Théroux, J.-F.; Lacaud, A.; Joganah, L.-E.; Lelievre, V.; Massotte, D.; Lutz, P.-E.; Gilsbach, R.; Salvat, E.; Barrot, M. A dual noradrenergic mechanism for the relief of neuropathic allodynia by the antidepressant drugs duloxetine and amitriptyline. J. Neurosci. 2018, 38, 9934-9954.
  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.
  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.
  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.
  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.
  219. Stark, P.; Fuller, R.W.; Wong, D.T. The pharmacologic profile of fluoxetine. J. Clin. Psychiatry 1985, 46, 7-13.
  220. Korzeniewska-Rybicka, I.; Płaźnik, A. Analgesic effect of antidepressant drugs. Pharmacol. Biochem. Behav. 1998, 59, 331-338.
  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.
  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.
  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.
  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.
  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.
  226. Baastrup, C.; Finnerup, N.B. Pharmacological management of neuropathic pain following spinal cord injury. CNS Drugs 2008, 22, 455-475.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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. Neuromolecular Med. 2014, 16, 308-321.
  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, U.S.A., 2011, pp. 565-582.
  239. Caruso, R.; Ostuzzi, G.; Turrini, G.; Ballette, F.; Recla, E.; DallʼOlio, 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.
  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.
  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.
  242. Bhana, N.; Goa, K.L.; McClellan, K.J. Dexmedetomidine. Drugs 2000, 59, 263-268.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  250. Tsutsui, Y.; Sunada, K. Adding dexmedetomidine to articaine increases the latency of thermal antinociception in rats. Anesth. Prog. 2020, 67, 72–78.
  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.
  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.
  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.
  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.
  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.
  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.
  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. Anest. & Inten. Care Med. 2019, 9, 555756.
  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.
  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.
  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.
  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.
  262. Vaida, G.T.; Moss, P.; Capan, L.M.; Turndorf, H. Prolongation of lidocaine spinal anesthesia with phenylephrine. Anesth. Analg. 1986, 65, 781-785.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  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.
  273. MacDonald, E.; Kobilka, B.K.; Scheinin, M. Gene targeting-homing in on α2-adrenoceptor-subtype function. Trends Pharmacol. Sci. 1997, 18, 211-219.
  274. Virtanen, R. Pharmacological profiles of medetomidine and its antagonist, atipamezole. Acta Vet. Scand. 1989, 85, 29-37.
  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.
  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.
  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.
  278. Slingsby, L.S.; Taylor, P.M. Thermal antinociception after dexmedetomidine administration in cats: a dose-finding study. J. Vet. Pharmacol. Therap. 2008, 31, 135-142.
  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.
  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.
  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.
  282. Julius, D. TRP channels and pain. Annu. Rev. Cell Dev. Biol. 2013, 29, 355-384.
  283. Yang, K; Kumamoto, E.; Furue, H.; Yoshimura, M. Capsaicin facilitates excitatory but not inhibitory synaptic transmission in substantia gelatinosa of the rat spinal cord. Neurosci. Lett. 1998, 255, 135-138.
  284. Kosugi, M.; Nakatsuka, T.; Fujita, T.; Kuroda, Y.; Kumamoto, E. Activation of TRPA1 channel facilitates excitatory synaptic transmission in substantia gelatinosa neurons of the adult rat spinal cord. J. Neurosci. 2007, 27, 4443-4451.
  285. Inoue, M.; Fujita, T.; Goto, M.; Kumamoto, E. Presynaptic enhancement by eugenol of spontaneous excitatory transmission in rat spinal substantia gelatinosa neurons is mediated by transient receptor potential A1 channels. Neuroscience 2012, 210, 403-415.
  286. Luo, Q.-T.; Fujita, T.; Jiang, C.-Y.; Kumamoto, E. Carvacrol presynaptically enhances spontaneous excitatory transmission and produces outward current in adult rat spinal substantia gelatinosa neurons. Brain Res. 2014, 1592, 44-54.
  287. Xu, Z.-H.; Wang, C.; Fujita, T.; Jiang, C.-Y.; Kumamoto, E. Action of thymol on spontaneous excitatory transmission in adult rat spinal substantia gelatinosa neurons. Neurosci. Lett. 2015, 606, 94-99.
  288. Kang, Q.; Jiang, C.-Y.; Fujita, T.; Kumamoto, E. Spontaneous L-glutamate release enhancement in rat substantia gelatinosa neurons by (-)-carvone and (+)-carvone which activate different types of TRP channel. Biochem. Biophys. Res. Commun. 2015, 459, 498-503.
  289. Jiang, C.-Y.; Wang, C.; Xu, N.-X.; Fujita, T.; Murata, Y.; Kumamoto, E. 1,8- and 1,4-cineole enhance spontaneous excitatory transmission by activating different types of transient receptor potential channels in the rat spinal substantia gelatinosa. J. Neurochem. 2016, 136, 764-777.
  290. 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. Phytomedicine Plus 2022, 2, 100155.
  291. 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.
  292. 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.
  293. Guimarães, A.G.; Quintans, J.S.S.; Quintans-Júnior, L.J. Monoterpenes with analgesic activity - a systematic review. Phytother. Res. 2013, 27, 1-15.
  294. Tsuchiya, H. Anesthetic agents of plant origin: a review of phytochemicals with anesthetic activity. Molecules 2017, 22, 1369.
  295. 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.
  296. 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.
  297. 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.
  298. 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.
  299. 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.
  300. 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.
  301. 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.
  302. 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.
  303. 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.
  304. 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.
  305. 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.
  306. 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.; Barbosa, R. Linalool blocks excitability in peripheral nerves and voltage-dependent Na+ current in dissociated dorsal root ganglia neurons. Eur. J. Pharmacol. 2010, 645, 86-93.
  307. 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.
  308. 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.
  309. Nozoe, T. Über die farbstoffe im holzteile des “hinoki” baumes. I. Hinokitin und hinokitiol. Bull. Chem. Soc. Jpn. 1936, 11, 295-298.
  310. Magori, N.; Fujita, T.; Kumamoto, E. Hinokitiol inhibits compound action potentials in the frog sciatic nerve. Eur. J. Pharmacol. 2018, 819, 254-260.
  311. 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. Invest. Dermatol. 1998, 110, 24-28.
  312. 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.
  313. 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.
  314. 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.
  315. 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.
  316. 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.
  317. 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.
  318. 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.
  319. 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.
  320. Doze, V.A.; Westphal, L.M.; White, P.F. Comparison of propofol with methohexital for outpatient anesthesia. Anesth. Analg. 1986, 65, 1189-1195.
  321. 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.
  322. Antkowiak, B.; Rammes, G. GABA(A) receptor-targeted drug development - new perspectives in perioperative anesthesia. Expert Opin. Drug Discov. 2019, 14, 683-699.
  323. 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.
  324. 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.
  325. 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.
  326. 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.
  327. Hijikata, Y. Analgesic treatment with Kampo prescription. Expert Rev. Neurother. 2006, 6, 795-802.
  328. 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.
  329. 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.
  330. Mochiki, E.; Yanai, M.; Ohno, T.; Kuwano, H. The effect of traditional Japanese medicine (Kampo) on gastrointestinal function. Surg. Today 2010, 40, 1105-1111.
  331. 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.; Benzie, I.F.F.; Wachtel-Galor, S., Eds.; CRC Press, Boca Raton, FL, U.S.A., 2011, Chapter 1.
  332. 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.
  333. 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.
  334. 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.
  335. 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.
  336. Nakamura, M.; Jang, I.-S. Indomethacin inhibits tetrodotoxin-resistant Na+ channels at acidic pH in rat nociceptive neurons. Neuropharmacology 2016, 105, 454-462.
  337. Joshi, S.K.; Mikusa, J.P.; Hernandez, G.; Baker, S.; Shieh, C.C.; Neelands, T.; Zhang, X.F.; Niforatos, W.; Kage, K.; Han, P.; Krafte, D.; Faltynek, C.; Sullivan, J.P.; Jarvis, M.F.; Honore, P. 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.
More
Video Production Service