Actions of Analgesics on Nerve Conduction: Comparison
Please note this is a comparison between Version 2 by Peter Tang and Version 1 by Eiichi Kumamoto.

Action potential (AP) conduction in nerve fibers plays an important role in transmitting nociceptive information from the periphery to cerebral cortex.  It is possible that analgesics depress nerve AP conduction, resulting in antinociception.  Many of analgesics are known to suppress nerve AP conduction and also voltage-sensitive Na+ and K+ channels involved in AP conduction. Compound action potential (CAP) has been used to know whether nerve AP conduction is affected by analgesics. This review article will introduce the effects of clinically-used analgesics, analgesic adjuvants and plant-derived analgesics on fast-conducting CAPs and voltage-sensitive Na+ and K+ channels involved in AP production. It is suggested that analgesics-induced inhibition of nerve AP conduction contributes to at least a part of their antinociceptive effects.

  • analgesic
  • antinociception
  • nerve conduction
  • sciatic nerve
  • compound action potential
  • Na+ channel
  • K+ channel

1. Introduction

Nociceptive information from the periphery to the cerebral cortex is mainly transmitted by action potential (AP) conduction in nerve fibers and chemical transmission at synapses (see [1,2][1][2] for reviews). Nociceptive pain is usually acute and relieved by narcotic analgesics such as opioids and antipyretic analgesics including non-steroidal anti-inflammatory drugs (NSAIDs). On the other hand, pain may persist or recur for longer than three months, which is called chronic pain. One type of chronic pain, neuropathic pain, which occurs as a result of damage to the peripheral or central nervous system (PNS and CNS, respectively), is characterized by a hyper-excitability of neurons near the injured neuronal tissues (see [3] for review). This type of pain is often resistant to analgesics such as opioids and NSAIDs and is thus treated by using analgesic adjuvants such as α2-adrenoceptor agonists, antiepileptics, antidepressants and local anesthetics (see [4,5,6,7,8,9,10,11][4][5][6][7][8][9][10][11] for reviews). Although the main target of analgesics and analgesic adjuvants, except for NSAIDs and local anesthetics, is generally synapses (see [12,13,14][12][13][14] for reviews), it is possible that all of their drugs inhibit nerve AP conduction, partly contributing to their inhibitory effects on pain.
The AP conduction is mediated by voltage-gated Na+ and K+ channels located in nerve fibers. Thus, a depolarizing stimulus given to a nerve fiber point activates Na+ channels expressed in membranes of the fiber, allowing Na+ entry to the cytoplasm, caused by the gradient of the electrochemical potential of Na+, leading to a self-regenerative production of AP. This in turn results in an outward current (membrane depolarization) in a fiber point adjacent to the point to produce opening of other Na+ channels and so forth. Such a production of AP subsides by a subsequent inactivation of Na+ channels and activation of K+ channels (see [15,16][15][16] for reviews).
In a bundle of nerve fibers exposed to insulator such as oil, sucrose or air, AP conduction in each fiber produces AP current flowing through nerve bundle surface having high resistance that can be measured as a potential difference, i.e., compound action potential (CAP), by using two electrodes put on the nerve. Sciatic nerve trunk dissected from frogs is a useful preparation to easily and stably record voltage-gated Na+-channel blocker tetrodotoxin (TTX)-sensitive and fast-conducting (possibly myelinated Aα-fiber mediated) CAPs by using the nerve trunk exposed to air (air-gap method). A voltage-gated delayed-rectifier K+-channel inhibitor tetraethylammonium increased the half-peak duration of the CAP with no change in its peak amplitude, indicating an involvement of K+ channels (see [17] for review). Although not only fast-conducting but also slow-conducting (C-fiber mediated) CAPs were recorded from the frog sciatic nerve, the latter had much smaller peak amplitude and conduction velocity than the former [18].

2. Actions of Analgesics on Nerve Conduction

2.1. Opioids

Opioids are well-known to inhibit glutamatergic excitatory transmission by activating opioid receptors in the CNS including the central terminals of primary-afferent fibers, resulting in antinociception ([28,29,30][19][20][21]; see [31,32][22][23] for reviews). Not only central but also peripheral terminal opioid receptors in primary-afferent neurons are thought to be involved in antinociception ([33,34,35,36][24][25][26][27]; see [37][28] for review). Opioids also exhibit a local anesthetic effect in the PNS. Although it has been reported in decerebrated cats that the perineural administration of an opioid morphine has no effect on CAPs in the superficial radial nerve [38][29], AP conduction in peripheral nerve fibers is generally blocked by opioids. For example, opioids such as fentanyl and sufentanil decreased the peak amplitudes of CAPs recorded from peripheral nerve fibers [39][30] and inhibited peripheral nerve AP conduction [40][31]. A morphine-induced CAP inhibition in mammalian peripheral nerve fibers was antagonized by a non-specific opioid-receptor antagonist naloxone, indicating an involvement of opioid receptors [41][32]. Consistent with this observation, binding and immunohistochemical studies have shown the localization of opioid receptors in mammalian peripheral nerve fibers [42,43,44][33][34][35]. It has been also demonstrated that a frog sciatic nerve CAP inhibition produced by opioids is sensitive to naloxone [45][36]. On the contrary, there are reports showing that opioids decrease CAP peak amplitudes [39][30] and suppress nerve AP conduction [40][31] in a manner resistant to naloxone.

2.1.1. Tramadol

The compound (1RS,2RS)-2-[(dimethylamino)methyl]-1-(3-methoxyphenyl)cyclohexanol hydrochloride (tramadol) is an orally-active opioid which is clinically used as an analgesic in the CNS [46][37]. Although tramadol is metabolized to various compounds such as mono-O-desmethyltramadol (M1) via N- and O-demethylation in animals and humans [47][38], M1 is a therapeutically active drug as a central analgesic [46][37]. One of cellular mechanisms for the antinociceptive effect of tramadol is the activation of μ-opioid receptors [48,49][39][40]. In agreement with this idea, M1 has the highest affinity for cloned μ-opioid receptors among the metabolites of tramadol. M1 is reported to inhibit glutamatergic excitatory transmission in spinal lamina II (substantia gelatinosa) neurons which play a crucial role in regulating nociceptive transmission to the spinal dorsal horn from the periphery, resulting in a decrease in the excitability of the neurons [50,51,52][41][42][43]. In addition to such a central activity, tramadol is known to have a local anesthetic effect following its intradermal injection in patients ([53,54,55][44][45][46]; see [56][47] for review). Consistent with this result, in vivo studies have shown a spinal somatosensory evoked potential inhibition produced by a direct application of tramadol to the rat sciatic nerve [57][48]. Tramadol reduced the peak amplitude of CAPs recorded from the frog sciatic nerve in a concentration-dependent manner in a range of 0.2 to 5 mM [19][49]. A similar CAP inhibitory action of tramadol has been reported by other investigators in the frog [58][50] and rat sciatic nerve [59,60][51][52]. Analysis based on the Hill equation demonstrated that half-maximal inhibitory concentration (IC50) value for tramadol to reduce frog sciatic nerve CAP amplitudes is 2.3 mM; this IC50 value was smaller by about three-fold than that (6.6 mM) reported previously for the frog sciatic nerve [58][50]. Rat sciatic nerve CAPs were inhibited by tramadol (37% peak amplitude reduction at 4 mM) less effectively than frog sciatic nerve’s ones [59][51]. This inhibitory action of tramadol in the frog sciatic nerve was not affected by the pretreatment of the sciatic nerve with naloxone (0.01 mM); a μ-opioid receptor agonist (D-Ala2, N-Me-Phe4, Gly5-ol)enkephalin (DAMGO; 1 μM) had no effect on frog sciatic nerve CAPs. Furthermore, CAPs were affected by much smaller extents by M1 (see below) that is similar in chemical structure to tramadol while exhibiting a higher affinity for μ-opioid receptors than tramadol [61][53]. These results indicate that the tramadol-induced CAP inhibition is not mediated by opioid receptors [19][49]. Consistent with this result, a spinal somatosensory evoked potential inhibition following the application of tramadol on rat sciatic nerves in vivo was resistant to naloxone [57][48]. Although tramadol inhibits noradrenaline (NA) and serotonin (5-hydroxytryptamine; 5-HT) reuptake at concentrations similar to those that activate μ-opioid receptors [62[54][55],63], a combination of inhibitors of the reuptake of NA and 5-HT (desipramine and fluoxetine, respectively; each 10 μM; see Section 3.3) did not affect frog sciatic nerve CAPs, indicating no involvement of NA and 5-HT reuptake inhibition in the CAP inhibition [19][49]. The CAP inhibition produced by tramadol is possibly due to an inhibition of voltage-gated Na+ and K+ channels involved in the production of AP. Tramadol concentration-dependently reduced the peak amplitude of TTX-sensitive Na+ channel currents recorded from dorsal root ganglion (DRG) neuroblastoma hybridoma cell line ND7/23 cells with an IC50 value of 0.194 mM [64][56] and from HEK293 cells expressing rat Nav1.2 channels with an IC50 value of 0.103 mM [65][57]. These values were smaller than that (2.3 mM) for frog sciatic nerve CAP inhibition. It has been demonstrated that tramadol suppresses the current amplitude of delayed rectifier K+-channels (Kv3.1a type) expressed in NG 108-15 cells with an IC50 of 0.025 mM, a value much less than 2.3 mM [66][58]. Such IC50 values for tramadol to inhibit CAPs, Na+ and K+ channels were higher than its clinically relevant concentration of about 2 μM in serum [52,67][43][59]. Unlike tramadol, M1 (1-2 mM) did not affect frog sciatic nerve CAPs. This was confirmed in the frog sciatic nerve whose CAPs were inhibited by tramadol (1 mM; [19][49]). CAP peak amplitudes were reduced by 9% by M1 at 5 mM. Consistent with such smaller effects of M1, APs conducting on rat primary-afferent fibers were not blocked when the effect of M1 (1 mM) on dorsal root-evoked excitatory postsynaptic currents was examined by applying the patch-clamp technique to lamina II neurons in spinal cord slices [51][42]. It is interesting to note that tramadol has -OCH3 bound to the benzene ring while M1 has -OH and thus that the methyl group is present in tramadol but not M1. This result indicates that the difference in chemical structure between tramadol and M1 is responsible for the distinction in CAP inhibition (see Figure 5a in [19][49] for a comparison of the chemical structures of the two compounds).

2.1.2. Other Opioids

In order to reveal whether such a structure-activity relationship is seen for other opioids, the effects of morphine, codeine and ethylmorphine on frog sciatic nerve CAPs were examined. Morphine concentration-dependently reduced CAP peak amplitude; this extent at 5 mM was 15%. Codeine, which has -OCH3 in place of -OH in morphine, at 5 mM reduced CAP peak amplitude by 30%. Moreover, CAPs were more effectively inhibited by ethylmorphine, where -OH of morphine is replaced by -OCH2CH3; amplitude reduction at 5 mM was 61%. IC50 value for ethylmorphine in reducing CAP peak amplitudes was 4.6 mM. CAP inhibitions produced by morphine (10 mM), codeine (5 mM) and ethylmorphine (2 mM) were resistant to naloxone (0.01 mM). Naloxone at a high concentration such as 1 mM by itself reduced by 9% CAP peak amplitudes, but did not affect morphine (10 mM) activity [20][60]. These results indicate no involvement of opioid receptors in the CAP inhibition produced by opioids, as reported previously in mammalian peripheral nerves [39,40,68][30][31][61]. A sequence of the opioid-induced CAP peak amplitude reduction was ethylmorphine > codeine > morphine. Thus, CAP amplitude reduction increased in extent with an increase in the number of -CH2 (see Figure 7A in [20][60] for a comparison of the chemical structures of the three opioids). This result is consistent with the above-mentioned observation that tramadol having -OCH3 in the benzene ring inhibits CAPs more effectively than M1 which is different from tramadol only in terms of the presence of -OH in the ring. Interestingly, this result was obtained in spite of the fact that the chemical structures of morphine, codeine and ethylmorphine are quite distinct from those of tramadol and M1 (see [69][62] for review). Since the increase in -CH2 number is thought to enhance lipophilicity of opioids, lipophilic opioid-channel interaction is suggested to play a pivotal role in nerve AP conduction block, as shown for local anesthetics [70,71][63][64]. This idea is supported by the observation that the potency in CAP inhibition in the rat sciatic nerve was in the order of isopropylcocaine (where the methyl ester group of cocaine is replaced with an isopropyl ester group) > cocaethylene (where its methyl ester group is replaced with an ethyl ester group) > cocaine [72][65]). It is interesting to note that the sequence of the affinity of opioids for μ-opioid receptors is morphine > codeine > ethylmorphine [73][66], the order of which is reversed to one for CAP inhibition. If the opioid-induced inhibition of CAPs is mediated by μ-opioid receptors, CAP inhibition sequence will be expected to be morphine > codeine > ethylmorphine. However, this sequence is not seen, a result being consistent with the idea that the opioids-induced frog sciatic nerve CAP inhibition is not mediated by opioid receptors. The same sequence as that in the frog sciatic nerve has been reported in the rat phrenic nerve [68][61], although there is a quantitative difference between the two studies. When compared at 5 mM, codeine-induced reduction (about 30%) in frog sciatic nerve CAP peak amplitude was much smaller than that (about 70%) in the rat phrenic nerve, while so a large distinction was not seen in morphine action (about 10%). Frog sciatic nerve CAPs were less sensitive to morphine than those in rabbit and guinea-pig vagus nerves in such that vagus nerve CAP peak amplitudes were reduced by 20–32% at 0.5 mM [41][32]. APs recorded intracellularly from rat DRG neurons having Aα/β myelinated primary-afferent fibers also exhibited the sequence of ethylmorphine > codeine ≥ morphine (IC50 = 0.70, 2.5 and 2.9 mM, respectively) in AP peak amplitude reduction; this inhibition was resistant to naloxone (0.01 mM) [74][67]. Although many drugs including narcotics, antiepileptics, local anesthetics, alcohols and barbiturates block AP conduction in peripheral nerve fibers, suggesting a nonspecific interaction of the drugs with membrane bilayers [75][68], the chemical structure-specific CAP inhibition produced by opioids indicates that opioids act on proteins such as voltage-gated Na+ and K+ channels (see [76][69] for review). Morphine is reported to suppress peak Na+ currents and steady-state K+ currents in single myelinated nerve fibers isolated from the frog sciatic nerve, leading to the prolongation of APs [77][70]. Intracellularly-applied morphine reduced voltage-gated Na+ and K+ channel current amplitudes in squid giant axons [78][71]. Bath-applied morphine reduced TTX-sensitive Na+ channel current amplitude in DRG neuroblastoma hybridoma cell line ND7/23 cells with an IC50 value of 0.378 mM [64][56], although Nav1.2 channels expressed in HEK293 cells were unaffected by morphine at 1 mM [65][57]. In support of such an idea about ion channel inhibition, it has been reported that an opioid meperidine, which is used for AP conduction blockade and thus analgesia, inhibits Na+-channels in a manner similar to that of lidocaine [79][72]. Table 1 summarizes IC50 values for frog sciatic nerve fast-conducting CAP inhibitions produced by opioids together with those for rat sciatic nerve CAPs and voltage-gated Na+ channels.
Table 1.
Comparison of IC
50
values in inhibiting frog or rat sciatic nerve fast-conducting CAPs and TTX-sensitive Na
+
channels among opioids.

Opioids

Frog CAP

Rat CAP

TTX-Sensitive

References

 

IC50 (mM)

IC50 (mM)

Na+ Channel Current

 
     

IC50 (mM)

 

Tramadol

2.3

37% reduction

0.194

[19,59,64][49][51][56]

With respect to other types of NSAIDs, salicylic acid-based (aspirin; 1 mM), propionic acid-based (ketoprofen, naproxen, ibuprofen, loxoprofen and flurbiprofen; each 1 mM) and enolic acid-based [meloxicam (0.5 mM) and piroxicam (1 mM)] NSAIDs had no effects on frog sciatic nerve CAP amplitudes [21][101]. CAP amplitude reductions produced by the NSAIDs would be mediated by an inhibition of TTX-sensitive voltage-gated Na+ channels that are involved in frog CAP production. In support of this idea, diclofenac decreased the peak amplitudes of TTX-sensitive Na+-channel currents in rat DRG [108][102] and mouse trigeminal ganglion neurons [109][103]. A similar diclofenac-induced Na+-channel inhibition has been reported in rat myoblasts [110][104] and ventricular cardiomyocytes [111][105]. Flufenamic acid as well as diclofenac decreased Na+-channel current amplitudes in rat hippocampal CA1 neurons [112,113,114][106][107][108]. Although IC50 value (0.22 mM) for flufenamic acid in frog sciatic nerve CAP inhibition was similar to that of Na+ channel inhibition (0.189 mM) in rat hippocampal CA1 neurons [114][108], IC50 value (0.94 mM) for diclofenac in CAP inhibition was much larger than those (0.00851 and 0.014 mM in rat myoblasts and DRG neurons, respectively) of Na+ channel inhibition [108,110][102][104]. Regarding rank order among NSAIDs, the order for CAP inhibition at 0.5 mM was flufenamic acid > diclofenac > indomethacin >> aspirin = naproxen = ibuprofen [21][101]; this was in part similar to those for Na+ channel inhibition in rat cardiomyocytes (diclofenac > naproxen ≥ ibuprofen; [111][105]) and also in rat DRG neurons (diclofenac > flufenamic acid > indomethacin > aspirin; [108][102]). With respect to TTX-resistant Na+ channels, diclofenac at 0.3 mM reduced peak current amplitudes by about 20% in rat trigeminal ganglion neurons [115][109]; Nav1.8 channel currents were inhibited by flufenamic acid and tolfenamic acid (current amplitude reduction: ca. 30 and 30%, respectively, at 0.1 mM; [116][110]). TTX-sensitive Nav1.7 channel currents were more sensitive to flufenamic acid and tolfenamic acid (reduction: ca. 60 and 70%, respectively, at 0.1 mM) than Nav1.8 ones [116][110]. Alternatively, chemical irritation-induced activity increase of cat corneal sensory nerve fibers was suppressed in extent by NSAIDs; this suppression was different in magnitude among distinct types of NSAIDs [109,117][103][111]. Na+-channel inhibition produced by NSAIDs appeared to be different in extent among preparations. Concentrations required for NSAIDs to have a significant inhibitory effect on frog sciatic nerve CAPs were in general higher than those needed to inhibit Na+ channels; this may be attributed to various reasons including the fact that not only Na+ channels but also K+ channels are involved in determining CAP amplitudes. To my knowledge, the effects of aceclofenac, indomethacin, etodolac, acemetacin, meclofenamic acid and mefenamic acid on voltage-gated Na+ channels have not been reported. Table 2 summarizes IC50 values for frog sciatic nerve fast-conducting CAP inhibitions produced by NSAIDs together with those for voltage-gated Na+ channels.
Table 2.
Comparison of IC
50
values in inhibiting frog sciatic nerve fast-conducting CAPs, TTX-sensitive or -resistant Na
+
channels among NSAIDs.

NSAIDs

Frog CAP

IC50 (mM)

TTX-Sensitive

Na+ Channel

Current IC50 (mM)

TTX-Resistant

Na+ Channel

Current IC50 (mM)

References

Acetic Acid-Based

Diclofenac

0.94

0.00851, 0.014

ca. 20% reduction

(0.3 mM)

[21,108,110,

   

(4 mM)

0.103

[65][57]

Mono-O-desmethyl

-tramadol

9% reduction

   

[19][49]

(5 mM)

     

Morphine

15% reduction

 

0.378

[20,64][56][60]

 

(5 mM)

     

Codeine

30% reduction

   

[20][60]

 

(5 mM)

     

Ethylmorphine

4.6

   

[20][60]

Here, where IC50 values are not available, it is partly shown for comparison how CAP amplitudes are reduced by drugs, where % value indicates the extent of the reduction at the concentration shown in parentheses. In clinical practice, although administration of opioids into the nerve sheath results in pain relief (for instance, see [80][73]), many of pain treatments by use of 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 [81][74] for review). It is possible that centrally-administrated opioids act on not only the CNS but also the PNS, because opioids are reported to be transported to the periphery from brain by P-glycoprotein [82][75]. In support of an important role of opioids in the PNS, subcutaneous administration of N-methyl-morphine, which did not pass through the blood brain barrier, resulted in antinociception in an acetic acid-writhing model in mice [35][26]. It has been reported that a subcutaneously-administrated opioid loperamide, which cannot penetrate into the brain, exhibited an antinociceptive effect in the formalin test in rats [34][25]. Such an action of opioids in the PNS appeared to be mediated by opioid receptors in peripheral terminals of primary-afferent fibers ([33,34,35,36,83][24][25][26][27][76]; see [37][28] for review). In addition, the inhibitory effect of opioids on nerve AP conduction also might contribute to local analgesia following the peripheral perineural administration of opioids (for instance, see [84][77]) that are expected to lead to a direct action of opioids at high doses on peripheral nerves. Since codeine is metabolized to morphine via O-demethylation in humans and animals ([85,86][78][79]; see [81][74] for review), peripherally-administrated codeine might have a similar effect to that of morphine.

2.2. NSAIDs

Antinociception produced by NSAIDs is mediated by various mechanisms such as (1) inhibition of the synthesis of prostaglandins from arachidonic acid by inhibiting the cyclooxygenase enzyme ([87,88][80][81]; see [89,90,91][82][83][84] for reviews), (2) inhibition of acid-sensitive ion channels [92][85] and transient receptor potential (TRP) channels [93,94][86][87], (3) activation of several K+ channels ([95,96,97,98,99][88][89][90][91][92]; see [100,101][93][94] for reviews), (4) substance P depletion [102][95], (5) an interaction with the adrenergic system [103][96] and (6) an involvement of opioids [104,105][97][98] and endocannabinoids (see [106][99] for review). The idea about an involvement of mechanisms other than cyclooxygenase inhibition in antinociception is supported by the observation that there is a dissociation between anti-inflammation and antinociception produced by NSAIDs [107][100]. An acetic acid-based NSAID diclofenac reduced frog sciatic nerve CAP peak amplitudes in a partially reversible manner. Diclofenac activity was concentration-dependent in a range of 0.01–1 mM with an IC50 value of 0.94 mM. Another acetic acid-based NSAID aceclofenac (a carboxymethyl ester of diclofenac) also exhibited a similar CAP inhibitory action. CAP peak amplitudes were concentration-dependently reduced by aceclofenac in a 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 had an efficacy smaller than those of 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 did so by 38%. Etodolac at 1 mM reduced CAP peak amplitudes by only 15%, and sulindac and felbinac at 1 mM had no effects on CAP peak amplitudes [21][101]. A similar frog sciatic nerve CAP inhibition was produced by fenamic acid-based NSAIDs (tolfenamic acid, meclofenamic acid, mefenamic acid and flufenamic acid) whose chemical structures are similar to those of diclofenac and aceclofenac. Tolfenamic acid concentration-dependently reduced CAP peak amplitudes in a range of 0.01–0.2 mM with an IC50 value of 0.29 mM. The activity of meclofenamic acid (where the chloro group bound to the benzene ring of tolfenamic acid is changed in number and position) was concentration-dependent in a range of 0.01–0.5 mM with an IC50 value of 0.19 mM. Moreover, mefenamic acid (where the chloro group bound to the benzene ring of tolfenamic acid is replaced by methyl group) concentration-dependently reduced CAP peak amplitudes in a range of 0.01–0.2 mM with the extent of 16% at 0.2 mM. CAP peak amplitudes were concentration-dependently reduced by 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) with an IC50 of 0.22 mM, a value comparable to those of tolfenamic acid and meclofenamic acid [21][101]. 2,6-Dichlorodiphenylamine and N-phenylanthranilic acid (which are similar in chemical structure to diclofenac and tolfenamic acid while being not NSAIDs) reduced frog sciatic nerve CAP peak amplitudes; the former compound lacks the -CH2COOH group of diclofenac and the latter one lacks chloro and methyl groups bound to the benzene ring of tolfenamic acid. 2,6-Dichloro- diphenylamine activity was concentration-dependent in a range of 0.001-0.1 mM with the extent of 45% at 0.1 mM; N-phenylanthranilic acid activity was concentration-dependent in a range of 0.01–2 mM with the extent of 23% at 1 mM [21][101].

115

]

[

101

]

[

102][104][109]

Aceclofenac

0.47

   

[21][101]

Indomethacin

38% reduction (1 mM)

   

[21][101]

Acemetacin

38% reduction (0.5 mM)

   

[21][101]

Etodolac

15% reduction (1 mM)

   

[21][101]

Sulindac

n.d.

(no effect, 1 mM)

   

[21][101]

Felbinac

n.d.

(no effect, 1 mM)

   

[21][101]

Fenamic Acid-Based

Tolfenamic acid

0.29

ca. 70% reduction (0.1 mM)

ca. 30% reduction (0.1 mM)

[21,116][101][110]

Meclofenamic acid

0.19

   

[21][101]

Mefenamic acid

16% reduction (0.2 mM)

   

[21][101]

Flufenamic acid

0.22

ca. 60% reduction

(0.1 mM)

ca. 30% reduction

(0.1 mM)

[21,[101116]][110]

   

0.189

 

[114][108]

Salicylic

Acid-Based

Aspirin

n.d.

(no effect, 1 mM)

   

[21][101]

Propionic

Acid-Based

Ketoprofen

n.d.

(no effect, 1 mM)

   

[21][101]

Naproxen

n.d.

(no effect, 1 mM)

   

[21][101]

Ibuprofen

n.d.

(no effect, 1 mM)

   

[21][101]

Loxoprofen

n.d.

(no effect, 1 mM)

   

[21][101]

Flurbiprofen

n.d.

(no effect, 1 mM)

   

[21][101]

Enolic

Acid-Based

Meloxicam

n.d.

(no effect, 0.5 mM)

   

[21][101]

Piroxicam

n.d.

(no effect, 1 mM)

   

[21][101]

Here, when IC50 values are not available, it is partly shown for comparison how the CAPs and channels are affected by drugs, where % value indicates the extent of the reduction at the concentration shown in parentheses; n.d.: not determined. NSAIDs (diclofenac, aceclofenac, tolfenamic acid, meclofenamic acid and flufenamic acid), which are more effective in frog sciatic nerve CAP inhibition compared to the other NSAIDs [21][101], have two benzene rings that bind a hydrophilic substituent group, both of which rings are linked by -NH- (see Figures 1Aa, 1Ba, 3Aa, 3Ba, 3Da in [21][101] for the chemical structures of the five NSAIDs), as seen in local anesthetics (see Section 3.4). Mefenamic acid (where one of the two benzene rings has a hydrophobic substituent group; see Figure 3Ca in [21][101]) appeared to be less effective, albeit not examined at a higher concentration due to a less solubility of this drug (see above). CAPs were effectively inhibited by 2,6-dichlorodiphenylamine and N-phenylanthranilic acid that are not NSAIDs while being similar in chemical structure to NSAIDs having two benzene rings (see Figures 4Aa and 4Ba in [21][101]). CAPs were also depressed by bisphenol A that have two benzene rings that bind a hydrophilic group such as -OH [26][112]. Much evidence demonstrates that the other actions of NSAIDs depend on their chemical structures. For instance, an involvement of NO-cGMP-K+ channels in antinociception mediated by NSAIDs was dependent on their chemical structures [98,118][91][113]. Nonselective cation channels in the rat exocrine pancreas were suppressed by flufenamic acid and mefenamic acid but not indomethacin, aspirin and ibuprofen [119][114]. There was a distinction in depressing TRP melastatin-3 channels between diclofenac and aceclofenac [94][87]. Although NSAIDs not only suppress but also activate TRP ankyrin-1 channels, this activation also differed in magnitude among NSAIDs [120][115]. Moreover, there was a distinction among NSAIDs in the activities of mitochondrial oxidative phosphorylation or electron transport system that may be involved in their adverse side effects [121][116]. Although the concentrations of NSAIDs tested in the frog sciatic nerve are generally much higher than those for voltage-gated Na+-channel inhibition, such high concentrations are likely when NSAIDs are used at high concentrations in the direct vicinity of nerve fibers. At least a part of analgesia caused by NSAIDs used as a dermatological drug for antinociception may be due to a nerve conduction inhibition through their inhibitory action on voltage-gated Na+ channels [122][117].

References

  1. Fields, H.L. Pain; McGraw-Hill: New York, NY, USA, 1987.
  2. Willis, W.D., Jr.; Coggeshall, R.E. Sensory Mechanisms of the Spinal Cord, 2nd ed.; Plenum: New York, NY, USA, 1991.
  3. Merskey, H. Clarifying definition of neuropathic pain. Pain 2002, 96, 408–409.
  4. 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.
  5. Finnerup, N.B.; Sindrup, S.H.; Jensen, T.S. The evidence for pharmacological treatment of neuropathic pain. Pain 2010, 150, 573–581.
  6. Jensen, T.S. Anticonvulsants in neuropathic pain: Rationale and clinical evidence. Eur. J. Pain 2002, 6 (Suppl. A), 61–68.
  7. Kamibayashi, T.; Maze, M. Clinical uses of α2-adrenergic agonists. Anesthesiology 2000, 93, 1345–1349.
  8. Lynch, M.E. Antidepressants as analgesics: A review of randomized controlled trials. J. Psychiatry Neurosci. 2001, 26, 30–36.
  9. 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.
  10. 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.
  11. 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.
  12. Fürst, S. Transmitters involved in antinociception in the spinal cord. Brain Res. Bull. 1999, 48, 129–141.
  13. 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.
  14. Zeilhofer, H.U.; Wildner, H.; Yévenes, G.E. Fast synaptic inhibition in spinal sensory processing and pain control. Physiol. Rev. 2012, 92, 193–235.
  15. 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.
  16. Levitan, I.B.; Karczmarek, L.K. The Neuron, 3rd ed.; Oxford University Press: New York, NY, USA, 2002.
  17. 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.
  18. Kobayashi, J.; Ohta, M.; Terada, Y. C fiber generates a slow Na+ spike in the frog sciatic nerve. Neurosci. Lett. 1993, 162, 93–96.
  19. 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.
  20. 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.
  21. Yoshimura, M.; North, R.A. Substantia gelatinosa neurones hyperpolarized in vitro by enkephalin. Nature 1983, 305, 529–530.
  22. North, R.A. Opioid actions on membrane ion channels. In Handbook of Experimental Pharmacology; Herz, A., Ed.; Springer: Berlin, Germany, 1993; Volume 104, pp. 773–797.
  23. Yaksh, T.L. Pharmacology and mechanisms of opioid analgesic activity. Acta Anaesthesiol. Scand. 1997, 41, 94–111.
  24. 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.
  25. 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.
  26. Smith, T.W.; Buchan, P.; Parsons, D.N.; Wilkinson, S. Peripheral antinociceptive effects of N-methyl morphine. Life Sci. 1982, 31, 1205–1208.
  27. Wenk, H.N.; Brederson, J.-D.; Honda, C.N. Morphine directly inhibits nociceptors in inflamed skin. J. Neurophysiol. 2006, 95, 2083–2097.
  28. Stein, C.; Schäfer, M.; Machelska, H. Attacking pain at its source: New perspectives on opioids. Nature Med. 2003, 9, 1003–1008.
  29. 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.
  30. 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.
  31. 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.
  32. Jurna, I.; Grossmann, W. The effect of morphine on mammalian nerve fibres. Eur. J. Pharmacol. 1977, 44, 339–348.
  33. Coggeshall, R.E.; Zhou, S.; Carlton, S.M. Opioid receptors on peripheral sensory axons. Brain Res. 1997, 764, 126–132.
  34. 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.
  35. Wenk, H.N.; Honda, C.N. Immunohistochemical localization of delta opioid receptors in peripheral tissues. J. Comp. Neurol. 1999, 408, 567–579.
  36. Hunter, E.G.; Frank, G.B. An opiate receptor on frog sciatic nerve axons. Can. J. Physiol. Pharmacol. 1979, 57, 1171–1174.
  37. Klotz, U. Tramadol—The impact of its pharmacokinetic and pharmacodynamic properties on the clinical management of pain. Arzneimittelforschung 2003, 53, 681–687.
  38. Lintz, W.; Erlacin, S.; Frankus, E.; Uragg, H. Metabolismus von Tramadol bei Mensch und Tier. Arzneimittelforschung 1981, 31, 1932–1943.
  39. Hennies, H.-H.; Friderichs, E.; Schneider, J. Receptor binding, analgesic and antitussive potency of tramadol and other selected opioids. Arzneimittelforschung 1988, 38, 877–880.
  40. 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.
  41. 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.
  42. Koga, A.; Fujita, T.; Piao, L.-H.; Nakatsuka, T.; Kumamoto, E. Inhibition by O-desmethyltramadol of glutamatergic excitatory transmission in adult rat spinal substantia gelatinosa neurons. Mol. Pain 2019, 15, 1744806918824243.
  43. 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.
  44. 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.
  45. 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.
  46. 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.
  47. Le Roux, P.J.; Coetzee, J.F. Tramadol today. Curr. Opin. Anaesth. 2000, 13, 457–461.
  48. 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.
  49. 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.
  50. 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.
  51. 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.
  52. 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.
  53. 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.
  54. 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.
  55. 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.
  56. 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.
  57. 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.
  58. 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.
  59. 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.
  60. 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.
  61. 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.
  62. Kumamoto, E.; Mizuta, K.; Fujita, T. Opioid actions in primary-afferent fibers—Involvement in analgesia and anesthesia. Pharmaceuticals 2011, 4, 343–365.
  63. 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.
  64. 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.
  65. 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.
  66. 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.
  67. 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.
  68. 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.
  69. Scholz, A. Mechanisms of (local) anaesthetics on voltage-gated sodium and other ion channels. Br. J. Anaesth. 2002, 89, 52–61.
  70. Hu, S.; Rubly, N. Effects of morphine on ionic currents in frog node of Ranvier. Eur. J. Pharmacol. 1983, 95, 185–192.
  71. 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.
  72. 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.
  73. 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.
  74. Gutstein, H.B.; Akil, H. Opioid analgesics. In Goodman & Gilman’s the Pharmacological Basis of Therapeutics, 11th ed.; Brunton, L.L., Lazo, J.S., Parker, K.L., Eds.; McGraw-Hill, Medical Publishing Division: New York, NY, USA, 2006; pp. 547–590.
  75. King, M.; Su, W.; Chang, A.; Zuckerman, A.; Pasternak, G.W. Transport of opioids from the brain to the periphery by P-glycoprotein: Peripheral actions of central drugs. Nat. Neurosci. 2001, 4, 268–274.
  76. 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.
  77. Mays, K.S.; Lipman, J.J.; Schnapp, M. Local analgesia without anesthesia using peripheral perineural morphine injections. Anesth. Analg. 1987, 66, 417–420.
  78. Cleary, J.; Mikus, G.; Somogyi, A.; Bochner, F. The influence of pharmacogenetics on opioid analgesia: Studies with codeine and oxycodone in the Sprague-Dawley/Dark Agouti rat model. J. Pharmacol. Exp. Ther. 1994, 271, 1528–1534.
  79. 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.
  80. Ferreira, S.H. Prostaglandins, aspirin-like drugs and analgesia. Nat. New Biol. 1972, 240, 200–203.
  81. 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.
  82. Grosser, T.; Smyth, E.; FitzGerald, G.A. Anti-inflammatory, antipyretic, and analgesic agents; pharmacotherapy of gout. In Goodman & Gilman’s the Pharmacological Basis of Therapeutics, 12th ed.; Brunton, L.L., Chabner, B.A., Knollmann, B.C., Eds.; McGraw-Hill, Medical Publishing Division: New York, NY, USA, 2011; pp. 959–1004.
  83. Simmons, D.L.; Botting, R.M.; Hla, T. Cyclooxygenase isozymes: The biology of prostaglandin synthesis and inhibition. Pharmacol. Rev. 1996, 56 (Suppl. 1), 387–437.
  84. Vane, J.R. Introduction: Mechanism of action of NSAIDs. Br. J. Rheumatol. 1996, 35 (Suppl. 1), 1–3.
  85. 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.
  86. 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.
  87. 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.
  88. Garg, P.; Sanguinetti, M.C. Structure-activity relationship of fenamates as Slo2.1 channel activators. Mol. Pharmacol. 2012, 82, 795–802.
  89. 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.
  90. 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.
  91. 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.
  92. 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.
  93. Guinamard, R.; Simard, C.; Del Negro, C. Flufenamic acid as an ion channel modulator. Pharmacol. Ther. 2013, 138, 272–284.
  94. 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.
  95. 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.
  96. Silva, L.C.R.; Castor, M.G.M.; Souza, T.C.; Duarte, I.D.G.; Romero, T.R.L. NSAIDs induce peripheral antinociception by interaction with the adrenergic system. Life Sci. 2015, 130, 7–11.
  97. 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.
  98. 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.
  99. Fowler, C.J. NSAIDs: eNdocannabinoid stimulating anti-inflammatory drugs? Trends Pharmacol. Sci. 2012, 33, 468–473.
  100. 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.
  101. 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.
  102. 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.
  103. Acosta, M.C.; Luna, C.; Graff, G.; Meseguer, V.M.; Viana, F.; Gallar, J.; Belmonte, C. Comparative effects of the nonsteroidal anti-inflammatory drug nepafenac on corneal sensory nerve fibers responding to chemical irritation. Investig. Ophthalmol. Vis. Sci. 2007, 48, 182–188.
  104. 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.
  105. Yarishkin, O.V.; Hwang, E.M.; Kim, D.; Yoo, J.C.; Kang, S.S.; Kim, D.R.; Shin, J.-H.-J.; Chung, H.-J.; Jeong, H.-S.; Kang, D.; et al. Diclofenac, a non-steroidal anti-inflammatory drug, inhibits L-type Ca2+ channels in neonatal rat ventricular cardiomyocytes. Korean J. Physiol. Pharmacol. 2009, 13, 437–442.
  106. 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.
  107. 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.
  108. 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.
  109. 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.
  110. 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.
  111. Chen, X.; Gallar, J.; Belmonte, C. Reduction by antiinflammatory drugs of the response of corneal sensory nerve fibers to chemical irritation. Investig. Ophthalmol. Vis. Sci. 1997, 38, 1944–1953.
  112. 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.
  113. 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.
  114. 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.
  115. 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.
  116. 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.
  117. 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.
More
Video Production Service