Next Article in Journal
Vitamin E and Silymarin Reduce Oxidative Tissue Damage during Gentamycin-Induced Nephrotoxicity
Next Article in Special Issue
Chronic Cannabigerol as an Effective Therapeutic for Cisplatin-Induced Neuropathic Pain
Previous Article in Journal
Supplementary Effects of Allium hookeri Extract on Glucose Tolerance in Prediabetic Subjects and C57BL/KsJ-db/db Mice
Previous Article in Special Issue
Pregabalin–Tolperisone Combination to Treat Neuropathic Pain: Improved Analgesia and Reduced Side Effects in Rats
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Exploring Cholinergic Compounds for Peripheral Neuropathic Pain Management: A Comprehensive Scoping Review of Rodent Model Studies

by
Edouard Montigné
1 and
David Balayssac
2,*
1
INSERM, U1107, NEURO-DOL, Université Clermont Auvergne, F-63000 Clermont-Ferrand, France
2
INSERM, U1107, NEURO-DOL, Université Clermont Auvergne, Direction de la Recherche Clinique et de l’Innovation, CHU Clermont-Ferrand, F-63000 Clermont-Ferrand, France
*
Author to whom correspondence should be addressed.
Pharmaceuticals 2023, 16(10), 1363; https://doi.org/10.3390/ph16101363
Submission received: 2 August 2023 / Revised: 1 September 2023 / Accepted: 14 September 2023 / Published: 27 September 2023
(This article belongs to the Special Issue Pharmacotherapy of Neuropathic Pain)

Abstract

:
Neuropathic pain affects about 7–8% of the population, and its management still poses challenges with unmet needs. Over the past decades, researchers have explored the cholinergic system (muscarinic and nicotinic acetylcholine receptors: mAChR and nAChR) and compounds targeting these receptors as potential analgesics for neuropathic pain management. This scoping review aims to provide an overview of studies on peripheral neuropathic pain (PNP) in rodent models, exploring compounds targeting cholinergic neurotransmission. The inclusion criteria were original articles on PNP in rodent models that explored the use of compounds directly targeting cholinergic neurotransmission and reported results of nociceptive behavioral assays. The literature search was performed in the PubMed and Web of Science databases (1 January 2000–22 April 2023). The selection process yielded 82 publications, encompassing 62 compounds. The most studied compounds were agonists of α4β2 nAChR and α7 nAChR, and antagonists of α9/α10 nAChR, along with those increasing acetylcholine and targeting mAChRs. Studies mainly reported antinociceptive effects in traumatic PNP models, and to a lesser extent, chemotherapy-induced neuropathy or diabetic models. These preclinical studies underscore the considerable potential of cholinergic compounds in the management of PNP, warranting the initiation of clinical trials.

1. Introduction

The International Association for the Study of Pain (IASP) defines pain as “an unpleasant sensory and emotional experience associated with, or resembling that associated with, actual or potential tissue damage”. In particular, peripheral neuropathic pain (PNP) is defined as “a direct consequence of a lesion or disease of the somatosensory nervous system and may be felt in areas with no tissue damage”, and more precisely, in the peripheral nervous system (PNS) [1]. These injuries of the peripheral somatosensory system result in maladaptive responses that can cause pain that is either spontaneous or evoked by sensory stimuli, as an increased response to a painful stimulus (hyperalgesia), or a painful response to a normally nonpainful stimulus (allodynia) [2].
Many etiologies can explain PNP, among which traumatic causes (e.g., phantom limb pain, postsurgical/traumatic neuropathy), toxic causes (e.g., chemotherapy-induced peripheral neuropathy—CIPN), and metabolic causes (e.g., hyperglycemic conditions) are found [3].
Approximately 7–8% of the general population worldwide experience PNP, with prevalence rates ranging from 0.9% to 17.9% [4]. Guidelines for the first-line pharmacological management of PNP include gabapentin, gabapentin extended release/enacarbil, pregabalin, serotonin and norepinephrine reuptake inhibitors (duloxetine, venlafaxine), and tricyclic antidepressants [5]. However, the efficacy of therapeutic strategies remains limited, with several and underassessed safety concerns [6,7,8], and innovative strategies are particularly needed [9].
The involvement of cholinergic neurotransmission in pain processes has been shown in many studies and could become a new therapeutic target for the management of PNP (for a review, see [10]).
Acetylcholine acts as one of the most prominent neurotransmitters in both the central nervous system (CNS) and PNS, particularly at the ganglionic level of the autonomous nervous system. After its release from cholinergic cells, acetylcholine is quickly metabolized by acetylcholine esterase (AChE) into choline and acetate, and thereafter, biosynthesized to acetylcholine by choline acetyltransferase (ChAT) at the neuronal level. Cholinergic neurotransmission involves two types of receptors: nicotinic acetylcholine receptors (nAChRs), and muscarinic acetylcholine receptors (mAChRs) [11].
The nAChRs are ionotropic receptors consisting of five individual subunits that arrange in a homo- or heteromeric combination and belong to the superfamily of pentameric ligand-gated ion channels (pLGICs), also known in vertebrates as the Cys-loop family. The member subunits of the Cys-loop family contain a disulfide cysteine bridge in the extracellular domain which closes a loop comprising 13 amino acids. Besides the nAChRs, the Cys-loop family includes the serotonin type 3 receptors (5-HT3), gamma aminobutyric acid type A receptors (GABAA) and glycine receptors [12]. A total of 19 different nAChR subunits have been identified: α1–α10, β1–β4, γ, δ, and ε subunits [12]. Myriad nAChR subtypes can be formed by different combinations of these subunits. For example, α4 and β2 subunits assemble together and comprise the most abundant nAChR subtype, α4β2 [13]. nAChRs regulate the flow of mainly sodium, potassium and calcium ions across the cell membrane [14].
The mAChRs are metabotropic receptors and are represented by five types of receptors, including m1AChR, m2AChR, m3AChR, m4AChR, and m5AChR. mAChRs are separated into two subtypes: excitatory (m1AChR, m3AChR, and m5AChR), and inhibitory (m2AChR and m4AChR). m1AChR, m3AChR, and m5AChR couple to a Gq protein, activating the phospholipase C pathway and increasing the intracellular calcium concentration. The m2AChRs and m4AChRs receptors couple to a Gi/o protein, inhibiting the adenylate cyclase pathway and resulting in a decrease in the amount of intracellular cyclic adenosine monophosphate [11,15].
The implications of the acetylcholine neurotransmission in pain have been demonstrated with a tonic cholinergic inhibition of spinal nociceptive transmission in rats [16], and nerve injury-induced loss of this cholinergic tone underlies the analgesic effects of cholinergic agonists in neuropathic pain [17,18]. Muscarinic signaling is vital in pain modulation, with direct activation of mAChRs reducing pain and mAChRs inhibition inducing nociceptive hypersensitivity [19]. The antinociceptive effects of systemically administered donepezil (AChE inhibitor) are linked to GABAergic signaling downstream of mAChRs, suggesting that cholinergic neurons’ inhibitory action may be mediated through GABA release [20]. Intrathecally administered AChE inhibitors, like neostigmine, reduce inflammatory hypersensitivity primarily through m2AChR activation in the spinal cord [21]. The m2AChR and m4AChR receptors predominantly mediate antinociception, while m1AChR and m3AChR have minimal involvement in this process [22].
At the spinal level, presynaptic cholinergic receptors play a significant role in modulating nociceptive transmission, with reports of both excitation and inhibition. Neuronal subpopulations in the dorsal root ganglia (DRG) express α7 nAChRs and respond to α7-selective positive allosteric modulators with a calcium rise, but the functional significance remains unclear [23]. mAChRs in the DRG and trigeminal ganglia modulate primary afferent input onto spinal or medullary dorsal horn neurons, leading to inhibition of glutamate release [24,25]. Activation of m2AChR and m4AchR inhibits excitatory synaptic inputs from primary afferents into the spinal cord by targeting voltage-gated calcium channels in primary sensory neurons [26]. Additionally, the m5AchR subtype has a more complex role, mediating positive effects on primary afferent terminals while potentially increasing glutamate release from spinal interneurons [26].
Supraspinal centers, such as the serotonergic raphe nuclei and adrenergic centers in the locus coeruleus and rostral ventromedial medulla (RVM), play a key role in endogenous nociception control through descending modulation of spinal function. Nicotinic cholinergic signaling in brainstem nuclei stimulates descending inhibitory pathways and mediates antinociceptive effects via interactions with α2 adrenergic, 5-HT1c/2, and 5-HT3 serotonergic receptors, and m2AchR in the lumbar spinal cord [27]. Studies also show antinociceptive effects from nAChR agonists in the RVM via activation of α4β2 nAChRs (and to a lesser extent, via α7 nAChRs) [28]. Finally, ChAT-Cre mice have demonstrated direct descending control of spinal sensory transmission by brainstem cholinergic neurons [29].
This scoping review aims to provide an overview of studies conducted in rodent PNP models exploring pharmacological compounds that target cholinergic neurotransmission, including the modulation of acetylcholine neurotransmission, nAChRs, and mAChRs.

2. Materials and Methods

We followed the Preferred Reporting Items for Systematic Review and Meta-Analysis extension for Scoping Reviews (PRISMA-ScR) guidelines in the study process [30].

2.1. Eligibility Criteria

The inclusion criteria encompassed original articles on PNP in rodent models (in vivo) that investigated the use of compounds directly targeting cholinergic neurotransmission and reported results of nociceptive behavioral assays.
The exclusion criteria encompassed non-English manuscripts, specific types of manuscripts (reviews, meta-analyses, case reports, books, and conference abstracts), publications lacking available abstracts, in vitro or ex vivo studies, and medicinal chemistry reports. Additionally, publications indirectly exploring cholinergic neurotransmission were excluded from the systematic review. For instance, studies using AChRs antagonists to investigate the involvement of AChRs in the analgesic effect of non-cholinergic compounds were not included.

2.2. Information Sources and Search Strategy

We conducted literature retrieval in the following databases: PubMed (MEDLINE, National Library of Medicine), and Web of Science (Clarivate Analytics PLC). No registers, websites, organizations, reference lists, or other sources were used. The bibliographical search was performed on 22 April 2023, and starting from 1 January 2000.
For PubMed, the sequence of keywords was “(acetylcholine) and ((neuropathy) or (neuropathic pain)) and ((rat) or (mouse)) and ((muscarinic) or (nicotinic))”. For Web of Science, the sequence of keywords was “acetylcholine (All Fields) AND (neuropathic pain) or (neuropathy) (All Fields) AND (rat) or (mouse) (All Fields) AND (muscarinic) or (nicotinic) (All Fields)”.

2.3. Study Selection, Data Collection Process and Data Items

All the references from PubMed and Web of Science were extracted and organized using Zotero software (version 6.0.26, Roy Rosenzweig Center for History and New Media) to create a Zotero bibliographic database. This database included the following details for each publication: authors, title, journal, year, abstract, and DOI (Digital Object Identifier). Subsequently, this Zotero bibliographic database was exported to Excel software (version 2021, Microsoft) for analysis.
Initial publication selection, based on title and abstract, was carried out by the authors EM and DB. Following this initial screening, all the authors conducted a second round of selection based on the full-text of the publications and in accordance with the inclusion/exclusion criteria. In cases where discrepancies regarding the inclusion/exclusion criteria arose for a publication, a consensus among the authors was sought to determine whether to include or exclude the publication.
The researchers extracted the following data from the included manuscripts: title, author, year, DOI, type of PNP, rodent species, sex of animals, name of the compounds, targeted AChRs, behavioral tests and main results.

3. Results

3.1. Study Selection and Characteristics

The process for selection and inclusion of publications is presented in Figure 1. A total of 82 publications were selected and analyzed, and these encompassed 62 cholinergic compounds.
Among the publications selected, 58 used animal models of traumatic PNP, 23 CIPN models and 5 diabetic PNP models (Supplementary Materials, Table S1).
For traumatic PNP, nine different models were used as follows:
-
Chronic constriction injury (CCI, n = 24);
-
Partial sciatic nerve ligation (PSL, n = 18);
-
Spinal nerve ligation (SNL, n = 13);
-
Spared nerve injury (SNI, n = 2);
-
Common peroneal nerve ligation (CPNL, n = 2);
-
Cuff model (CM, n = 1);
-
Sciatic nerve crush injury (SCNI, n = 1);
-
Sciatic nerve transection (SNT, n = 1);
-
Tibial nerve transection (TNT, n = 1).
Among the CIPN models, the oxaliplatin model (n = 14) was the most used, but other models such as paclitaxel (n = 6), vincristine (n = 4), and bortezomib (n = 1) were also described. Finally, two models of diabetic PNP were used: streptozotocin-induced peripheral neuropathy (SZT, n = 4), and the high-fat diet (HFD, n = 1).

3.2. Pharmacological Compounds Increasing the Acetylcholine Neurotransmission

To modulate acetylcholine neurotransmission, seven pharmacological compounds with different mechanisms of actions have been tested, including AChE inhibitors, inhibitors of acetylcholine exocytosis, and acetylcholine precursors (Table 1). All these compounds act to increase the quantity of acetylcholine in the synaptic cleft.
AChE inhibitors were the most commonly investigated compounds, such as ambenonium chloride, donepezil, huperzine-A, neostigmine, and physostigmine. All these compounds have been tested in the three animal models of PNP. Most of the compounds induced antinociceptive effects, both on mechanical and thermal hypersensitivity [20,31,32,33,34,35,36,37]. A low dose of neostigmine (intrathecal injection, 0.3 ng) did not produce significant antinociceptive effects [38].
Botulinum neurotoxin type A has been used to inhibit the exocytosis of acetylcholine and has shown a persistent antinociceptive effect on mechanical hypersensitivity [39,40].
Finally, citicoline, a precursor of choline, exhibited antinociceptive properties on mechanical hypersensitivity after intracerebroventricular injection or nerve application [41,42,43].
Table 1. Description of studies assessing pharmacological compounds modulating the acetylcholine neurotransmission.
Table 1. Description of studies assessing pharmacological compounds modulating the acetylcholine neurotransmission.
CompoundsTargetsDoses, RoutesModelsSpecies
(Sex)
Behavioral
Assays
EffectsRef.
Ambenonium
chloride
AChE
inhibitors
0.05 mg/kg, i.p.TraumaticMouse (♂)Von Frey hairs
Conditioned place preference
↘ M
0 spontaneous pain
[34]
Donepezil0.3–1 mg/kg, i.p.TraumaticRat (♂)Paw pressure↘ M[20]
0.3–1.0 mg/kg, i.p.TraumaticRat (♂)Paw pressureDose-dependent ↘ M
(0.6 and 1 mg/kg)
[31]
5–10 mg/kg, p.o.CIPN +
traumatic
Rat (♂)Electronic Von Frey
Paw immersion 10 °C and 46 °C
↘ M + T[32]
5 mg/kg, p.o.CIPNRat (♂)Electronic Von Frey
Tail immersion 10 °C
↘ M + T[33]
Huperzine-A0.1–0.15 mg/kg, i.p.TraumaticMouse (♂)Von Frey hairs
Conditioned place preference
↘ M
0 spontaneous pain
[34]
Neostigmine2 mg, i.t.TraumaticRat (♂)Von Frey hairs↘ M[36]
0.3–3 ng, i.t.TraumaticMouse (♂)Plantar test
Von Frey hairs
Dose-dependant ↘ M + T
(3 ng)
[38]
0.1–0.5 µg, i.t.DiabeticRat (♂)Von Frey hairsDose-dependent ↘ M
(0.5 µg)
[37]
Physostigmine15 nmol, i.t.TraumaticMouse (♂)Von Frey hairs↘ M[35]
BoNT/AAcetylcholine
exocytosis
inhibitors
15 pg, i.p.TraumaticMouse (♂)Electronic Von Frey↘ M[39]
BoNT/A Dysport®
BoNT/A Botox®
20 U/kg, s.p.CIPNRat (♂)Paw pressure↘ M[40]
CiticolineAcetylcholine
precursor
0.4–0.8 mL (100 µmol/L), p.n.TraumaticRat (♂)Von Frey hairs↘ M[41]
0.5–2 µmol, i.c.v.TraumaticRat (♂)Paw pressureDose- and time-dependent ↘ M
(1 and 2 µmol)
[42]
0.5–2 µmol, i.c.v.CIPNRat (♂)Paw pressureDose- and time-dependent ↘ M
(1 and 2 µmol)
[43]
BoNT/A: botulinum neurotoxin type A; AChE: acetylcholine esterase; p.o.: per os; i.p.: intraperitoneal; i.t.: intratechal; p.n.: perinervous; s.p.: subplantar; i.c.v.: intracerebroventricular; ↘: decrease; 0: no effect; M: mechanical sensitivity; T: thermal sensitivity; ♀: female; ♂: male.

3.3. Pharmacological Compounds Targeting Nicotinic Acetylcholine Receptors

3.3.1. Unspecific Targeting of Nicotinic Acetylcholine Receptors

Four compounds unspecifically targeting nAChRs were identified: epibatidine, nicotine, S(−)-nornicotine, and R(+)-nornicotine (Table 2).
Nicotine was the most studied compound at various doses and routes (1–30 nmol i.t., 0.1–2 mg/kg s.c., and 0.3–1.75 mg/kg i.p.), mainly in traumatic PNP models, with one study on a CIPN model (paclitaxel). In several publications, nicotine demonstrated analgesic properties on mechanical and thermal hypersensitivity. It showed a decrease in mechanical allodynia in a time- and dose-dependent manner [18,44,45,46,47,48,49]. However, chronic administration of nicotine (24 mg/kg/day, osmotic pump) induced a dose-dependent and stable mechanical hypersensitivity [50]. Similarly, repeated administrations of nicotine induced a decrease in its analgesic effects (heat hypersensitivity) [51]. Other studies have also reported a reduction in mechanical and thermal hypersensitivity [52,53]. Two enantiomers, S(−)-nornicotine and R(+)-nornicotine, dose-dependently reversed mechanical hypersensitivity [54].
Finally, epibatidine has also been found to dose-dependently reverse mechanical allodynia in traumatic PNP [18,48,55].
Table 2. Description of studies assessing pharmacological compounds targeting nicotinic acetylcholine receptors.
Table 2. Description of studies assessing pharmacological compounds targeting nicotinic acetylcholine receptors.
CompoundsTargetsDoses, RoutesModelsSpecies
(Sex)
Behavioral
Assays
EffectsRef.
EpibatidinenAChRs
agonists
0.03–0.3 nmol, i.t.TraumaticMouse (♂)Electronic Von Frey
Plantar test
Dose-dependent ↘ M
(all doses effective)
[18]
0.3–10.0 µg/kg, s.c.TraumaticRat (♂)Paw pressureDose-dependent ↘ M
(all doses effective)
[55]
0.036–0.36 pmol, i.t.TraumaticRat (♂)Paw pressureDose-dependent ↘ M[48]
Nicotine0.1–1.5 mg/kg, s.c.TraumaticMouse (♂)Von Frey hairsDose- and time-dependent↘ M
(1 and 1.5 mg/kg)
[45]
0.1–10 nmol, i.t.TraumaticRat (♂)Electronic Von FreyDose- and time-dependent↘ M
(1 nmol)
[46]
3–30 nmol, i.t.TraumaticMouse (♂)Electronic Von Frey
Plantar test
Dose-dependent ↘ M
(10 and 30 nmol)
[18]
2.2–6.5 nmol, i.t.TraumaticRat (♂)Paw pressureDose-dependent ↘ M[48]
0.25–1.75 mg/kg, i.p.TraumaticMouse (♂ + ♀)Von Frey hairsDose-dependent ↘ M[49]
0.25–25 µg, i.c.v.TraumaticMouse (♂ + ♀)Von Frey hairsDose-dependent ↘ M[49]
0.25–17.5 µg, i.t.TraumaticMouse (♂ + ♀)Von Frey hairsDose-dependent ↘ M[49]
25–100 µg, i.pl.TraumaticMouse (♂ + ♀)Von Frey hairsDose-dependent ↘ M[49]
4 or 10 mg/kg/day, s.c.TraumaticRat (♂)Paw pressure↗ M[50]
20 nmol/day/4 days, p.n.
1–20 nmol, p.n.
TraumaticMouse (♂)Von Frey hairs
Plantar test
↘ M (preventive effect)
Dose-dependent ↘ M + T
(5 and 20 nmol)
[47]
2 mg/kg, s.c.TraumaticMouse (♂)Von Frey hair↘ T[51]
1 mg/kg, i.v.TraumaticMouse (♂)Von Frey hairs
Plantar test
Tail-flick
↘ M + T[52]
20 nmol, p.n.TraumaticMouse (♂)Von Frey hair
Plantar test
↘ M + T[53]
0.3–0.9 mg/kg, i.p.
24 mg/kg/day, s.c.
CIPNMouse (♂)Von Frey hairsDose-dependent ↘ M
(0.6 and 0.9 mg/kg)
↘ M (preventive effect)
[44]
S(-)-nornicotine5–20 mg/kg, i.p.TraumaticRat (♂)Paw pressureDose-dependent ↘ M
(20 mg/kg)
[54]
R(+)-nornicotine10–15 mg/kg, i.p.TraumaticRat (♂)Paw pressureDose-dependent ↘ M
(15 mg/kg)
[54]
i.p.: intraperitoneal; i.t.: intratechal; s.c.: subcutaneous; i.c.v.: intracerebroventricular; i.pl.: intraplantar; p.n.: perinervous; ↘: decrease; ↗: increase; M: mechanical sensitivity; T: thermal sensitivity; ♀: female; ♂: male.

3.3.2. Specific Targeting of α4β2 Nicotinic Acetylcholine Receptors

Ten compounds specifically targeting of α4β2 nAChRs have been identified. Nine of them were agonists ([123/125I]5IA, A-366833, A-85380, NS9283, ABT-418, ABT-594, Bee venom, C-9515, RJR-2403, and TC-2559 [46,55,56,57,58,59,60,61,62,63,64,65,66]), and one a partial agonist (Sazetidine A [56]) (Table 3). These compounds were primarily tested in traumatic PNP models.
All the tested compounds produced antinociceptive effects, and mostly on mechanical hypersensitivity. However, NS9283 and A-85380 (administered at low pmol dose intrathecally) did not show any effects on mechanical or thermal hypersensitivity in traumatic PNP [48,67].
Table 3. Description of studies assessing pharmacological compounds targeting α4β2 nicotinic acetylcholine receptors.
Table 3. Description of studies assessing pharmacological compounds targeting α4β2 nicotinic acetylcholine receptors.
CompoundsTargetsDoses, RoutesModelsSpecies
(Sex)
Behavioral
Assays
EffectsRef.
[123/125I]5IAα4β2
nAChR
agonists
1–100 nmol, i.c.v.TraumaticRat (♂)Von Frey hairsDose-dependent ↘ M
(50 and 100 nmol)
[57]
1–10 nmol, i.c.v.TraumaticRat (♂)Von Frey hairsDose-dependent ↘ M
(3 and 10 nmol)
[63]
A-3668331.9–19 µmol/kg, i.p.TraumaticRat (♂)Von Frey hairsDose-dependent ↘ M
(all doses effective)
[64]
1–6 mg/kg, i.p.Traumatic +
Diabetic +CIPN
Rat (♂)Paw pressure↘ M[65]
A-853800.125–1 µmol/kg, i.p.TraumaticRat (♂)Von Frey hairsDose-dependent ↘ M
(0.75 and 1 µmol/kg
[66]
<pmol, i.t.TraumaticRat (♂)Paw pressure0 M[48]
NS928335 mmol/kg, i.p.TraumaticRat (♂)Von Frey hair0 M[67]
ABT-4181–20 nmol, p.n.TraumaticMouse (♂)Von Frey hairs
Plantar test
↘ M + T[53]
ABT-5943–100 µg/kg, s.c.TraumaticRat (♂)Von Frey hairsDose-dependent ↘ M
(10–100 µg/kg)
[55]
0.01–0.3 µmol/kg, i.p.CIPNRat (♂)Von Frey hairsDose-dependent ↘ M
(0.1–0.3 µmol/kg)
[59]
Bee venom0.25 mg/kg, s.c.CIPNRat (♂)Tail immersion test↘ T[61]
C-95150.001–0.01 mg/kg, i.p.TraumaticRat (♂)Von Frey hairsDose-dependent ↘ M
(0.003 and 0.01 mg/kg)
[58]
RJR-24031–100 nmol, i.t.TraumaticRat (♂)Electronic Von Frey↘ M[46]
TC-25592.28–22.8 µmol/kg, s.c.
20 nmol, p.n.
TraumaticMouse (♂)Von Frey hairsDose-dependent ↘ M
(22.8 µmol/kg)
↘ M
[56]
1–20 nmol, p.n.TraumaticMouse (♂)Von Frey hairs
Plantar test
↘ M + T[53]
0.3–3 mg/kg, i.p.TraumaticRat (♂)Von Frey hairsDose-dependent ↘ M
(1 and 3 mg/kg)
[62]
20 nmol, p.n.
10 mg/kg, s.c.
DiabeticMouse (♂)Von Frey hair↘ M[60]
Sazetidine Aα4β2
nAChR
partial agonist
0.2–20 nmol, p.n.TraumaticMouse (♂)Von Frey hairsDose-dependent ↘ M
(20 nmol)
[56]
i.p.: intraperitoneal; i.t.: intratechal; s.c.: subcutaneous; i.c.v.: intracerebroventricular; p.n.: perinervous; s.c.: subcutaneous; ↘: decrease; 0: no effect; M: mechanical hypersensitivity; T: thermal hypersensitivity; ♀: female; ♂: male.

3.3.3. Specific Targeting of α7 Nicotinic Acetylcholine Receptors

Seventeen compounds targeting α7 nAChRs were identified. Among them, ten were α7 nAChR agonists, two were α7 nAChR silent agonists, three were positives allosteric modulators (PAM) of α7 nAChRs, one was an α3β4 and α7 nAChRs antagonist, and one was an α7 nAChRs + m4AChRs agonist (Table 4). These compounds were mostly tested in traumatic PNP models.
Among the α7 nAChR agonists, all the tested compounds ((R)-ICH3, choline, cobratoxin, DM489, GAT107, PAM-4, PHA-543613, PNU-28298, PNU-282987, and TC-7020) exhibited antinociceptive effects on mechanical and thermal hypersensitivity [46,68,69,70,71,72,73,74,75,76]. While GAT107 was effective on mechanical hypersensitivity, no effect was reported on thermal hypersensitivity [72].
The two α7 nAChR silent agonists, NS6740 and R-47, demonstrated antinociceptive effects on mechanical hypersensitivity [77,78].
Regarding the three PAM of α7 nAChRs, only PAM-2 and PNU-120596 had analgesic effects [69,73,79], but NS1738 had no effect on mechanical and thermal hypersensitivity [69].
The α3β4 and α7 nAChRs antagonist, α-conopeptides Eu1.6 [80], and the α7 nAChRs + m4AChRs agonist, DXL-A-24 [81], both decreased mechanical hypersensitivity.
Table 4. Description of studies assessing pharmacological compounds targeting α7 nicotinic acetylcholine receptor.
Table 4. Description of studies assessing pharmacological compounds targeting α7 nicotinic acetylcholine receptor.
CompoundsTargetsDoses, RoutesModelsSpecies
(Sex)
Behavioral
Assays
EffectsRef.
(R)-ICH3α7
nAChR
agonist
30 mg/kg, p.o.CIPNRat (♂)Paw pressure
Von Frey hairs
Cold plate (4 °C)
↘ M + T[70]
Choline100 nmol, i.t.TraumaticRat (♂)Electronic Von Frey↘ M[46]
Cobratoxin0.56–4.5 µg/kg, i.t.TraumaticRat (♂)Paw pressure
Tail-flick
Dose-dependent ↘ M + T
(1.12 and 4.5 µg/kg)
[68]
DM4893–10 mg/kg, p.o.DiabeticMouse (♂)Cold plate 4 °C↘ T[73]
10–30 mg/kg, p.o.CIPNMouse (♂)Cold plate 4 °C↘ T[73]
GAT1071–10 mg/kg, i.p.TraumaticMouse (♂)Von Frey hair
Plantar test
Dose- and time-dependent ↘ M
(3 and 10 mg/kg)
↘ T
[72]
PAM-41–2 mg/kg, i.p.TraumaticMouse (♂)Von Frey hairsDose- and time-dependent ↘ M
(2 mg/kg)
[74]
PHA-54361312 μg, i.t.TraumaticRat (♂)Electronic von frey
Plantar test
↘ M + T[71]
1–6 mg/kg, s.c.TraumaticMouse (♂)Von Frey hairsDose-dependent ↘ M
(6 mg/kg)
[69]
PNU-282981–30 mg kg, p.o.TraumaticRat (♂)Paw pressureDose-dependent ↘ M[75]
PNU-28298730 mg/kg, p.o.CIPNRat (♂)Paw pressure↘ M[70]
1 µg/kg, i.t.TraumaticRat (♂)Paw pressure
Tail-flick
↘ M + T[68]
TC-70201–10 mg/kg/day, s.c.TraumaticRat (♂)Von Frey hairs↘ M[76]
NS6740α7
nAChR
silent agonist
1–9 mg/kg, i.p.TraumaticMouse (♂)Von Frey hairsDose-dependent ↘ M
(9 mg/kg)
[77]
R-470–10 mg/kg, p.o.CIPNMouse (♂)Von Frey hairsDose- and time-dependent ↘ M
(5 and 10 mg/kg)
[78]
PAM-2α7
nAChR
PAM
3 mg/kg, p.o.Diabetic + CIPNMouse (♂)Cold plate 4 °C↘ T[73]
2–8 mg/kg, i.p.TraumaticMouse (♂)Von Frey hairsDose-dependent ↘ M
(6 and 8 mg/kg)
[79]
NS173830 mg/kg, i.p.TraumaticMouse (♂)Von Frey hairs
Plantar test
0 M + T[69]
PNU-1205961–8 mg/kg, i.p.TraumaticMouse (♂)Von Frey hairs
Plantar test
↘ M + T[69]
α-conopeptides Eu1.6α3β4/α7
nAChR
antagonist
0.5–24.9 μg/kg, i.m.TraumaticRat (♂)Paw pressure↘ M[80]
DXL-A-24α7 nAChR/
m4AChR
agonist
0.25–1 mg/kg, p.o.
daily
TraumaticRat (♂)Von Frey hairs
Hot plate 50 °C
Dose- and time-dependent ↘ M + T
(0.5 and 1 mg/kg)
[81]
PAM: positive allosteric modulator; p.o.: per os; i.p.: intraperitoneal; i.m.: intramuscular; i.t.: intratechal; s.c.: subcutaneous; ↘: decrease; 0: no effect; M: mechanical sensitivity; T: thermal sensitivity; ♀: female; ♂: male.

3.3.4. Specific Targeting of α9/α10 Nicotinic Acetylcholine Receptors

We identified seventeen α9/α10 nAChR antagonists (Table 5). These compounds have been tested in traumatic and CIPN PNP models, but not in diabetic models.
Most of the studies explored α-conotoxins and derivatives as potent antagonists of α9/α10 nAChR, including α-conotoxin AuIB, α-conotoxin MII, α-conotoxin Mr1.1 [S4Dap], α-conotoxin RgIA, α-conotoxin RgIA4, α-conotoxin RgIA-5474, [2,8]-alkyne Vc1.1 3, GeXIVA[1,2], GeXIVA[1,4], α-conotoxin Vc1.1, Vc1.1[N9R], and [P6O]Vc.1.1. Doses ranged from 0.036 to 60 µg (i.m. route), from 0.128 to 80 µg/kg (s.c. route), and from 0.02 to 2 nmol (i.t. route).
Nearly all the α9/α10 nAChR antagonists, including (±)-18-MC, (+)-catharanthine, α-conotoxins AuIB, α-conotoxins MII, α-conotoxins Mr1.1 [S4Dap], α-Conotoxin RgIA, α-conotoxins RgIA4, α-Conotoxin RgIA-5474, [2,8]-alkyne Vc1.1 3, α-conotoxins Vc1.1, Vc1.1[N9R], GeXIVA[1,2], GeXIVA[1,4], ZZ-204G, and ZZ1-61c, presented antinociceptive effects on mechanical and thermal hypersensitivity [80,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104]. [P6O]Vc.1.1 had no effect on mechanical hypersensitivity [85], while α-conotoxins RgIA4 (assessed in paclitaxel-related CIPN model) [98], GeXIVA[1,2] (assessed in oxaliplatin-related CIPN model) [102], and ZZ-204G (assessed with the tail flick test in naïve animals) [89], had no effects on thermal hypersensitivity.
Table 5. Description of studies assessing pharmacological compounds targeting α9/α10 nicotinic acetylcholine receptors.
Table 5. Description of studies assessing pharmacological compounds targeting α9/α10 nicotinic acetylcholine receptors.
CompoundsTargetsDoses, RoutesModelsSpecies
(Sex)
Behavioral
Assays
EffectsRef.
(±)-18-MCα9/α10
nAChR
antagonist
72 mg/kg, p.o.CIPNMouse (♂)Cold plate 4 °C↘ T[90]
(+)-catharanthine36–72 mg/kg, p.o.CIPNMouse (♂)Cold plate 4 °C↘ T[90]
α-conotoxin AuIB0.02–2 nmol, i.t.TraumaticRat (♂)Von Frey hairsDose-dependent ↘ M
(2 nmol)
[82]
0.36–36 µg, i.m.TraumaticRat (♂)Von Frey hairs↘ M[83]
α-conotoxin MII0.02–2 nmol, i.t.TraumaticRat (♂)Von Frey hairsDose-dependent ↘ M
(2 nmol)
[82]
0.36–36 µg, i.m.TraumaticRat (♂)Von Frey hairs↘ M[83]
α-conotoxin Mr1.1 [S4Dap]0.5–25 µg/kg, nrTraumaticRat (♂)Electronic Von FreyDose-dependent ↘ M
(25 µg/kg)
[91]
α-conotoxin RgIA2–10 nmol, i.m.TraumaticRat (♂)Paw pressure
Electronic Von Frey
↘ M[84]
α-conotoxin RgIA40.02–0.2 nmol, i.m.TraumaticRat (♂)Paw pressureDose-dependent ↘ M
(0.2 nmol)
[92]
100 mg/kg, i.m.CIPNMouse (♂)Hot plate 47 °C
Electronic Von Frey
Acetone test
↘ T + M[93]
2–10 nmol, i.m.CIPNRat (♂)Cold plate 4 °C↘ T[94]
0.128–80 µg/kg, s.c.CIPNMouse + Rat (♂)Paw pressure
Cold plate 4 °C
Dose-dependent ↘ M + T
(all doses effective)
[95]
40 µg/kg, s.c.CIPNMouse (nr)Cold plate
(decrease of 10 °C/min)
↘ T[96]
40 µg/kg, s.c.CIPNMouse (nr)Cold plate
(decrease of 10 °C/min)
↘ T[97]
16–80 µg/kg, s.c.CIPNRat (♂)Von Frey hairs
Cold plate 5 °C
Plantar test
Time-dependent ↘ M
0 T
[98]
α-conotoxin RgIA-54744–40 µg/kg, s.c.CIPNMouse (nr)Cold plate
(decrease of 10 °C/min)
↘ T[99]
α-conotoxin Vc1.10.02–2 nmol, i.t.TraumaticRat (♂)Von Frey hairsDose-dependent ↘ M
(2 nmol)
[82]
0.36–36 µg, i.m.TraumaticRat (♂)Von Frey hairsDose-dependent ↘ M
(all doses effective)
[83]
0.36–3.6 µg, i.m.TraumaticRat (♂)Paw pressureDose-dependent ↘ M
(3.6 µg)
[101]
0.036–0.36 µg, i.m.TraumaticRat (♂)Paw pressureDose-dependent ↘ M[92]
60 µg, i.m.TraumaticRat (♂)Von Frey hairs↘ M[85]
27.2–54.2 μg/kg, i.m.TraumaticRat (♂)Paw pressure↘ M[80]
[2,8]-alkyne Vc1.1 30.03–0.1 mg/kg, i.m.TraumaticRat (♂)Von Frey hairsDose-dependent ↘ M
(0.1 mg/kg)
[100]
Vc1.1[N9R]0.3–15 nmol/kg, i.m.TraumaticRat (♂)Paw pressure↘ M[86]
[P6O]Vc.1.160 µg, i.m.TraumaticRat (♂)Von Frey hairs0 M[85]
GeXIVA[1,2]0.5–2 nmol, i.m.TraumaticRat (♂)Electronic Von FreyDose-dependent ↘ M
(All doses effective)
[104]
0.3–1.2 nmol, i.m.TraumaticRat (♂)Electronic Von Frey
Von Frey hairs
↘ M[87]
32–128 nmol/kg, i.m.CIPNRat (♂)Von Frey hairs
+ acetone test + tail-flick
Dose-dependent ↘
(128 nmol/kg)
0 T
[102]
0.45 mg/kg, i.m.CIPNRat (♂)Von Frey hairs
Tail-flick
↘ M + T[103]
GeXIVA[1,4]0.5–2 nmol, i.m.TraumaticRat (♂)Electronic Von FreyDose-dependent ↘ M
(1 and 2 nmol)
[104]
Oligoarginine R820 mg/kg, i.m.CIPNMouse (♂)Hot plate 47 °C
Electronic Von Frey
Acetone test
↘ M + T[93]
ZZ-204G3.6–3600 µg/kg, i.p.TraumaticRat (♂)Paw pressure
Tail-flick
Dose-dependent ↘ M
(360 and 3600 µg/kg)
0 T
[89]
ZZ1-61c100 µg/kg/day, i.p.CIPNRat (♂)Paw pressure↘ M[88]
p.o.: per os; i.p.: intraperitoneal; i.m.: intramuscular; i.t.: intratechal; s.c.: subcutaneous; nr: not reported; ↘: decrease; 0: no effect; M: mechanical hypersensitivity; T: thermal hypersensitivity; ♀: female; ♂: male.

3.4. Pharmacological Compounds Targeting Muscarinic Acetylcholine Receptors

Eight compounds targeting mAChRs have been identified, including one mAChR non-selective agonist (oxotremorine), one mAChR antagonist (scopolamine), one agonist of m2/m4AChRs and antagonist of m1/m3/m5AChRs (PTAC), one m4/α7AChRs agonist (DXL-A-24), one m1/m4AChRs agonist (McNA-343), one m1AChR agonist (PPBI), and two m1AChR antagonists (pirenzepine and VU0255035) (Table 6). These compounds were mostly tested in traumatic PNP models.
All these compounds induced antinociceptive effects in PNP models and mostly on mechanical hypersensitivity [38,81,105,106,107,108,109,110]. Only McNA-343 (m1/m4AChRs agonist) did not alleviate thermal hypersensitivity [38]. Scopolamine (mAChRs antagonist) injected in the anterior cingulate cortex decreased the autotomy behavior of animals (sciatic nerve section) [107].
Surprisingly, both agonist (PPBI) and antagonists (pirenzepine and VU0255035) of m1AChRs demonstrated antinociceptive effects in PNP models [109,110].
Table 6. Description of studies assessing pharmacological compounds targeting muscarinic acetylcholine receptors.
Table 6. Description of studies assessing pharmacological compounds targeting muscarinic acetylcholine receptors.
CompoundsTargetsDoses, RoutesModelsSpecies
(Sex)
Behavioral
Assays
EffectsRef.
OxotremorinemAChRs
non-selective agonist
5–10 μg, i.t.TraumaticRat (♂)Von Frey hairs
Ethyl chloride spray
Plantar test
↘ M + T[105]
PTACm2-4 AChRs
agonist
m1-3-5
AChRs
antagonist
0.05–0.1 mg/kg, i.p.TraumaticMouse (♂)Von Frey hairs↘ M[106]
ScopolaminemAChRs
antagonist
0.4 μg/μL, ACCTraumaticRat (♂)Daily autotomy scores↘ autotomy[107]
DXL-A-24m4AChR
+ α7 nAChR
agonist
0.25–1 mg/kg, p.o.TraumaticRat (♂)Von Frey hairsDose-dependent ↘ M
(0.5 and 1 mg/kg)
[81]
McNA-343m1/m4AChRs
agonist
18.9–1890 pmol, ACCTraumaticRat (♂)Electronic Von FreyDose-dependent ↘ M
(189 and 1890 pmol)
[108]
3–10 μg, i.t.TraumaticMouse (♂)Plantar test
Von Frey hairs
Paw pressure
↘ M
0 T
[38]
PPBIm1AChR
agonist
0.2–100 mmol/kg, p.o.Traumatic
+ CIPN
Mouse +
Rat (♂)
Tail immersion 13 °C
Acetone test
Plantar test
Dose-dependent
↘ M + T
[109]
Pirenzepinem1AChR
antagonist
10 mg/kg/d, s.c.Diabetic
+ CIPN
Mouse +
Rat (♂ + ♀)
Von Frey hairs↘ M[110]
VU025503510 mg/kg, i.p.DiabeticMouse +
Rat (♂ + ♀)
Von Frey hairs↘ M[110]
i.p.: intraperitoneal; i.t.: intratechal; s.c.: subcutaneous; p.o.: per os; ACC: anterior cingulate cortex injection; ↘: decrease; 0: no effect; M: mechanical sensitivity; T: thermal sensitivity; ♀: female; ♂: male.

4. Discussion

This review aimed to assess the state of the art regarding compounds targeting acetylcholine neurotransmission for the management of PNP in rodent models.
Among the 82 selected publications, 62 cholinergic compounds were assessed in rodent models of PNP. Almost all the compounds had antinociceptive effects in the PNP models, demonstrating the interest in targeting acetylcholine neurotransmission for pain management. Most of the studies assessed the compounds using traumatic models of PNP, and to a lesser extent, in CIPN or diabetic models. Two third of the studies were conducted with rats, and one third with mice. Most of the studies utilized male animals, and only two studies employed both male and female animals. However, sex differences exist in the context of PNP, particularly concerning the sexually dimorphic nature of neuroimmune pathophysiology [111], which highlights the importance of including subjects of both sexes in preclinical pain research [112].
In two-thirds of the assays, compounds were administered via systemic routes, with the i.p. route accounting for 23.2%, the i.m. route for 18.8%, the s.c. route for 13.6%, and the p.o. route for 11.6%. The remaining third of the assays employed local routes of administration, with the most representative ones being the i.t. route for 18.8%, p.n. for 5.4%, and the i.c.v. route for 4.5%. In spite of the interesting use of local administrations for defining precise mechanisms of action, systemic routes should be encouraged to improve the translationality of the results and drugability of the compounds. The assessment of the analgesic effects of compounds was primarily conducted using tests exploring mechanical hypersensitivity, which accounted for approximately 75% of the tests. Thermal hypersensitivity was less explored, comprising approximately 25% of the tests. Interestingly, in some cases, the same compound demonstrated effectiveness on mechanical hypersensitivity but not on thermal hypersensitivity (e.g., the α7 nAChR agonist GAT107 [72]; the α9/α10 nAChR antagonists α-conotoxin RgIA4 [98], GeXIVA[1,2] [84], and ZZ-204G [89]; and the m1/m4AChRs agonist McNA-343 [38]). Therefore, the exploration of the analgesic activity of these compounds should be encouraged on both modalities (mechanical and thermal) to better define their effects.
The present review highlights that research has primarily focused on nAChRs compared with mAChRs (Figure 2). Among the nAChRs, the α4β2 nAChR and α7 nAChR agonists, and α9/α10 nAChR antagonists were the main compounds explored. The first approach was driven by a global activation of nAChRs, thanks to nicotine [18,44,45,46,47,48,49,50,51,52,53]. Epibatidine, a toxic alkaloid isolated and identified from Epipedobates tricolor skin (for a review see [113]), was also used in a similar approach [18,48,55], and even though this compound is recognized today as an α4β2 nAChR agonist, it is also able to target α7 nAChR and α3β4 nAChR [113]. A more specific targeting of nAChRs has been tested with specific agonists of α4β2 nAChR and α7 nAChR, and antagonists of α9/α10 nAChR. The latter were mainly focused on α-conotoxins and derivatives originating from venomous marine cone snails (Conus species) [114].
α4β2 nAChR stands out as the most prevalent heteromeric subtype within the brain. Typically labeled as α4β2* nAChR, the asterisk acknowledges the potential involvement of additional subunits. When presented in isolation, the α4 and β2 subunits coalesce to form two distinctive functional isoforms: the low-sensitivity (α4)3(β2)2 and high-sensitivity (α4)2(β2)3 nAChRs. Notably, the lower acetylcholine sensitivity of (α4)3(β2)2 nAChR constitutes the majority within the cortex [115]. This α4β2* nAChR variant assumes a crucial role in nicotine dependence and represents a pharmacological target for smoking cessation aids like varenicline [115].
In contrast, the homopentameric α7 nAChR lacks the synaptic functional adaptations observed in muscle-type and ganglionic nAChRs. The absence of non-alpha subunits influences the agonist binding site, shifting high-affinity agonist binding toward desensitized states and enhancing sensitivity to the widely available acetylcholine precursor, choline [116]. Unique cellular and subcellular distribution patterns for α7 nAChR indicate its distinct roles. These distinctive patterns encompass widespread expression in non-neuronal cells, including those within the immune system. Notably, α7 nAChR plays a distinctive role in regulating the cholinergic anti-inflammatory pathway. An intriguing attribute of open α7 nAChR is its heightened calcium permeability [116]. Although this characteristic could potentially lead to excitotoxicity akin to NMDA-type glutamate receptors, the normally low open probability of the α7 nAChR receptor channel offsets this risk [116]. It is noteworthy that modifications in intracellular calcium concentration following α7 nAChR stimulation predominantly arise from the release of calcium stored intracellularly rather than calcium influx through α7 nAChR channels, implying a metabotropic-like effect [116].
α9/α10 nAChR represents a more recently studied subtype. Its primary functions have been elucidated in inner ear mechanosensory hair cells. Considered a non-neuronal nAChR, the α9/α10 subtype demonstrates negligible expression of α9 and very low expression of α10 subunits in murine DRG [12,117]. However, in human DRG, around 25% of sensory neurons have been found to express α9 and α10 mRNA [117]. Notably, the α9/α10 nAChR holds potential as an intriguing target for inner ear disorders from a pharmacotherapeutic perspective [12]. Beyond auditory disorders, blocking α9/α10 nAChRs has displayed analgesic effects linked to the modulation of immune cells and inflammatory processes [13].
Excluding α9/α10 nAChR, nAChRs are distributed across neurons in peripheral nociceptive nerve fibers, DRG, the dorsal horn of the spinal cord, the RVM in the brainstem, and the periaqueductal gray, all of which play pivotal roles in pain perception and processing [118]. Moreover, α4β2 nAChRs have been identified within astrocytes in the spinal cord and regions of the brain associated with pain processing [119]. α7 nAChRs display expression in immune and non-immune cells responsible for cytokine production, exercising control over immune cell functions, suppressing the production of pro-inflammatory cytokines, and mitigating inflammatory processes [116].
Finally, the targeting of mAChRs was clearly less explored, with few compounds and a widespread targeting of both agonists and antagonists for specific and unspecific mAChRs (Figure 2). No specific mAChRs emerged as clear pharmacological targets. Moreover, the definition of the compounds’ activities can be obscure. For example, both agonists and antagonists of the same m1AChRs induced antinociceptive effects [109,110], which raises doubt about the selectivity of these compounds. This targeting of mAChRs for pain management remains clearly underexplored, and more research is need.
Another interesting point is that compounds modulating the acetylcholine neurotransmission (e.g., AChE inhibitors, acetylcholine exocytosis inhibitors, and acetylcholine precursors), which ultimately increase the quantity of acetylcholine in the synaptic cleft, demonstrated analgesic effects (Figure 2) [20,31,32,33,34,35,36,37,39,40,41,42,43]. This unspecific strategy stands in stark contrast to the targeting of specific AChRs for pain modulation described in this review. By increasing the quantity of acetylcholine, these compounds promote global activation of nAChRs and mAChRs, which can be opposite to the sought effect in the case of α9/α10 nAChRs, where an inhibition of the latter is needed for an analgesic effect [120].
Studies on nicotine have explored the desensitization of its activity after chronic exposure, showing a decrease in its antinociceptive effects and even nociceptive sensitization, mediated by a desensitization of α4β2 nAChRs [50]. This desensitization was related to an increase in the phosphorylation of cAMP response element-binding protein (pCREB) in the spinal cord, which is highly correlated with the degree of mechanical hypersensitivity [50]. It is well demonstrated that tobacco smoking has acute antinociceptive effects [121] and chronic pronociceptive ones [122]. Tobacco smoking is a risk factor for chronic pain [122]. This desensitization of nAChRs has also been reported for α7 nAChR and α3β4 nAChR, driving the research from agonists of specific nAChRs to positive allosteric modulators of these nAChRs. Positive allosteric modulators can increase the binding affinity and/or efficacy of an orthosteric agonist (endogenous or exogenous) [123].
To our knowledge, besides BoNT/A, which is mainly authorized for various spasticity disorders, no cholinergic compound is currently authorized for the clinical management of PNP. The use of cholinergic compounds in patients experiencing pain could raise safety concerns regarding the cholinergic mechanism of action. We can extrapolate adverse effects of these compounds from those of AChE inhibitors used in Alzheimer’s disease treatment and medications used for smoking cessation (nicotine replacement therapy and varenicline). The most frequently reported adverse events of AChE inhibitors are gastrointestinal issues, including nausea, vomiting, diarrhea, and anorexia [124]. In the case of nicotine replacement therapy, the most commonly reported adverse events include headache, dizziness/light-headedness, nausea/vomiting, gastrointestinal symptoms, sleep/dream problems, and cardiovascular effects (palpitations, chest pain) [125]. With varenicline, the most frequently reported adverse events are nausea, insomnia, abnormal dreams, and headache [126].

Limitations of This Review

The aim of this review was to present an overview of the current studies focused on compounds that modulate cholinergic neurotransmission in the management of PNP in rodent models. This aim was not to report effective compounds on PNP. However, we believe that there is still a publication bias related to the fact that many negative results have not been published [127], and consequently, several studies focused on compounds that modulate cholinergic neurotransmission would be underreported.

5. Conclusions

The objective of this scoping review was to identify the primary targets and compounds studied in animal models of PNP. Besides mAChRs, nAChRs remain the most extensively investigated targets, with specific compounds such as α4β2 nAChR and α7 nAChR agonists, as well as α9/α10 nAChR antagonists. Notably, for the latter group, a majority of studies have focused on conotoxins and derivatives, representing innovative compounds. In contrast, when it comes to mAChRs, no distinct cholinergic target can be definitively defined based on this review.
The next steps involve a clear delineation of the antinociceptive mechanisms of action, including a more precise identification of the targeted AChRs and the cellular pathways involved, which, for many of these cholinergic targets, appear to encompass both neuronal and non-neuronal elements, including the immune system.
Overall, these preclinical studies underscore the considerable potential of cholinergic compounds in the management of PNP, warranting the initiation of clinical trials.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ph16101363/s1. Table S1. Description of animal models of peripheral neuropathic pain.

Author Contributions

Conceptualization, D.B.; methodology, E.M. and D.B.; software, E.M. and D.B.; validation, E.M. and D.B.; formal analysis, E.M. and D.B.; investigation, E.M. and D.B.; resources, D.B.; data curation, E.M. and D.B.; writing—original draft preparation, E.M. and D.B.; writing—review and editing, E.M. and D.B.; visualization, E.M. and D.B.; supervision, D.B.; project administration, D.B.; funding acquisition, D.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by International Research Centre 3: European Centre for Health and Human Mobility—Clermont Auvergne University. The APC was funded by CHU Clermont-Ferrand.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing is not applicable.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Raja, S.N.; Carr, D.B.; Cohen, M.; Finnerup, N.B.; Flor, H.; Gibson, S.; Keefe, F.J.; Mogil, J.S.; Ringkamp, M.; Sluka, K.A.; et al. The Revised International Association for the Study of Pain Definition of Pain: Concepts, Challenges, and Compromises. Pain 2020, 161, 1976–1982. [Google Scholar] [CrossRef] [PubMed]
  2. Murphy, D.; Lester, D.; Clay Smither, F.; Balakhanlou, E. Peripheral Neuropathic Pain. NeuroRehabilitation 2020, 47, 265–283. [Google Scholar] [CrossRef] [PubMed]
  3. Meacham, K.; Shepherd, A.; Mohapatra, D.P.; Haroutounian, S. Neuropathic Pain: Central vs. Peripheral Mechanisms. Curr. Pain Headache Rep. 2017, 21, 28. [Google Scholar] [CrossRef]
  4. van Hecke, O.; Austin, S.K.; Khan, R.A.; Smith, B.H.; Torrance, N. Neuropathic Pain in the General Population: A Systematic Review of Epidemiological Studies. Pain 2014, 155, 654–662. [Google Scholar] [CrossRef]
  5. Finnerup, N.B.; Attal, N.; Haroutounian, S.; McNicol, E.; Baron, R.; Dworkin, R.H.; Gilron, I.; Haanpää, M.; Hansson, P.; Jensen, T.S.; et al. Pharmacotherapy for Neuropathic Pain in Adults: A Systematic Review and Meta-Analysis. Lancet Neurol. 2015, 14, 162–173. [Google Scholar] [CrossRef]
  6. Cuménal, M.; Selvy, M.; Kerckhove, N.; Bertin, C.; Morez, M.; Courteix, C.; Busserolles, J.; Balayssac, D. The Safety of Medications Used to Treat Peripheral Neuropathic Pain, Part 2 (Opioids, Cannabinoids and Other Drugs): Review of Double-Blind, Placebo-Controlled, Randomized Clinical Trials. Expert Opin. Drug Saf. 2021, 20, 51–68. [Google Scholar] [CrossRef]
  7. Selvy, M.; Cuménal, M.; Kerckhove, N.; Courteix, C.; Busserolles, J.; Balayssac, D. The Safety of Medications Used to Treat Peripheral Neuropathic Pain, Part 1 (Antidepressants and Antiepileptics): Review of Double-Blind, Placebo-Controlled, Randomized Clinical Trials. Expert Opin. Drug Saf. 2020, 19, 707–733. [Google Scholar] [CrossRef]
  8. Finnerup, N.B.; Kuner, R.; Jensen, T.S. Neuropathic Pain: From Mechanisms to Treatment. Physiol. Rev. 2021, 101, 259–301. [Google Scholar] [CrossRef] [PubMed]
  9. Torrance, N.; Ferguson, J.A.; Afolabi, E.; Bennett, M.I.; Serpell, M.G.; Dunn, K.M.; Smith, B.H. Neuropathic Pain in the Community: More under-Treated than Refractory? Pain 2013, 154, 690–699. [Google Scholar] [CrossRef]
  10. Naser, P.V.; Kuner, R. Molecular, Cellular and Circuit Basis of Cholinergic Modulation of Pain. Neuroscience 2018, 387, 135–148. [Google Scholar] [CrossRef]
  11. Colangelo, C.; Shichkova, P.; Keller, D.; Markram, H.; Ramaswamy, S. Cellular, Synaptic and Network Effects of Acetylcholine in the Neocortex. Front. Neural Circuits 2019, 13, 24. [Google Scholar] [CrossRef]
  12. Elgoyhen, A.B. The A9α10 Acetylcholine Receptor: A Non-Neuronal Nicotinic Receptor. Pharmacol. Res. 2023, 190, 106735. [Google Scholar] [CrossRef]
  13. Hone, A.J.; McIntosh, J.M. Nicotinic Acetylcholine Receptors: Therapeutic Targets for Novel Ligands to Treat Pain and Inflammation. Pharmacol. Res. 2023, 190, 106715. [Google Scholar] [CrossRef] [PubMed]
  14. Ho, T.N.T.; Abraham, N.; Lewis, R.J. Structure-Function of Neuronal Nicotinic Acetylcholine Receptor Inhibitors Derived From Natural Toxins. Front. Neurosci. 2020, 14, 609005. [Google Scholar] [CrossRef]
  15. Picciotto, M.R.; Higley, M.J.; Mineur, Y.S. Acetylcholine as a Neuromodulator: Cholinergic Signaling Shapes Nervous System Function and Behavior. Neuron 2012, 76, 116–129. [Google Scholar] [CrossRef] [PubMed]
  16. Zhuo, M.; Gebhart, G.F. Tonic Cholinergic Inhibition of Spinal Mechanical Transmission. Pain 1991, 46, 211–222. [Google Scholar] [CrossRef] [PubMed]
  17. Matsumoto, M.; Xie, W.; Inoue, M.; Ueda, H. Evidence for the Tonic Inhibition of Spinal Pain by Nicotinic Cholinergic Transmission through Primary Afferents. Mol. Pain 2007, 3. [Google Scholar] [CrossRef]
  18. Rashid, M.H.; Ueda, H. Neuropathy-Specific Analgesic Action of Intrathecal Nicotinic Agonists and Its Spinal GABA-Mediated Mechanism. Brain Res. 2002, 953, 53–62. [Google Scholar] [CrossRef]
  19. Fiorino, D.F.; Garcia-Guzman, M. Muscarinic Pain Pharmacology: Realizing the Promise of Novel Analgesics by Overcoming Old Challenges. In Muscarinic Receptors; Fryer, A.D., Christopoulos, A., Nathanson, N.M., Eds.; Handbook of Experimental Pharmacology; Springer: Berlin/Heidelberg, Germany, 2012; Volume 208, pp. 191–221. ISBN 978-3-642-23273-2. [Google Scholar]
  20. Kimura, M.; Hayashida, K.; Eisenach, J.C.; Saito, S.; Obata, H. Relief of Hypersensitivity after Nerve Injury from Systemic Donepezil Involves Spinal Cholinergic and γ-Aminobutyric Acid Mechanisms. Anesthesiology 2013, 118, 173–180. [Google Scholar] [CrossRef]
  21. Yoon, S.; Kwon, Y.; Kim, H.; Roh, D.; Kang, S.; Kim, C.; Han, H.; Kim, K.; Yang, I.; Beitz, A. Intrathecal Neostigmine Reduces the Zymosan-Induced Inflammatory Response in a Mouse Air Pouch Model via Adrenomedullary Activity: Involvement of Spinal Muscarinic Type 2 Receptors. Neuropharmacology 2005, 49, 275–282. [Google Scholar] [CrossRef]
  22. Cai, Y.; Chen, S.; Han, H.; Sood, A.K.; Lopez-Berestein, G.; Pan, H. Role of M2, M3, and M4 Muscarinic Receptor Subtypes in the Spinal Cholinergic Control of Nociception Revealed Using siRNA in Rats. J. Neurochem. 2009, 111, 1000–1010. [Google Scholar] [CrossRef] [PubMed]
  23. Shelukhina, I.; Paddenberg, R.; Kummer, W.; Tsetlin, V. Functional Expression and Axonal Transport of A7 nAChRs by Peptidergic Nociceptors of Rat Dorsal Root Ganglion. Brain Struct. Funct. 2015, 220, 1885–1899. [Google Scholar] [CrossRef] [PubMed]
  24. Jeong, S.-G.; Choi, I.-S.; Cho, J.-H.; Jang, I.-S. Cholinergic Modulation of Primary Afferent Glutamatergic Transmission in Rat Medullary Dorsal Horn Neurons. Neuropharmacology 2013, 75, 295–303. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, H.-M.; Chen, S.-R.; Pan, H.-L. Regulation of Glutamate Release From Primary Afferents and Interneurons in the Spinal Cord by Muscarinic Receptor Subtypes. J. Neurophysiol. 2007, 97, 102–109. [Google Scholar] [CrossRef] [PubMed]
  26. Chen, S.-R.; Chen, H.; Yuan, W.-X.; Wess, J.; Pan, H.-L. Differential Regulation of Primary Afferent Input to Spinal Cord by Muscarinic Receptor Subtypes Delineated Using Knockout Mice. J. Biol. Chem. 2014, 289, 14321–14330. [Google Scholar] [CrossRef]
  27. Iwamoto, E.T.; Marion, L. Adrenergic, Serotonergic and Cholinergic Components of Nicotinic Antinociception in Rats. J. Pharmacol. Exp. Ther. 1993, 265, 777–789. [Google Scholar]
  28. Jareczek, F.J.; White, S.R.; Hammond, D.L. Plasticity in Brainstem Mechanisms of Pain Modulation by Nicotinic Acetylcholine Receptors in the Rat. eNeuro 2017, 4, 1–16. [Google Scholar] [CrossRef]
  29. Stornetta, R.L.; Macon, C.J.; Nguyen, T.M.; Coates, M.B.; Guyenet, P.G. Cholinergic Neurons in the Mouse Rostral Ventrolateral Medulla Target Sensory Afferent Areas. Brain Struct. Funct. 2013, 218, 455–475. [Google Scholar] [CrossRef] [PubMed]
  30. Tricco, A.C.; Lillie, E.; Zarin, W.; O’Brien, K.K.; Colquhoun, H.; Levac, D.; Moher, D.; Peters, M.D.J.; Horsley, T.; Weeks, L.; et al. PRISMA Extension for Scoping Reviews (PRISMA-ScR): Checklist and Explanation. Ann. Intern. Med. 2018, 169, 467–473. [Google Scholar] [CrossRef]
  31. Kimura, M.; Saito, S.; Obata, H. Dexmedetomidine Decreases Hyperalgesia in Neuropathic Pain by Increasing Acetylcholine in the Spinal Cord. Neurosci. Lett. 2012, 529, 70–74. [Google Scholar] [CrossRef] [PubMed]
  32. Ferrier, J.; Bayet-Robert, M.; Dalmann, R.; El Guerrab, A.; Aissouni, Y.; Graveron-Demilly, D.; Chalus, M.; Pinguet, J.; Eschalier, A.; Richard, D.; et al. Cholinergic Neurotransmission in the Posterior Insular Cortex Is Altered in Preclinical Models of Neuropathic Pain: Key Role of Muscarinic M2 Receptors in Donepezil-Induced Antinociception. J. Neurosci. 2015, 35, 16418–16430. [Google Scholar] [CrossRef]
  33. Selvy, M.; Mattévi, C.; Dalbos, C.; Aissouni, Y.; Chapuy, E.; Martin, P.-Y.; Collin, A.; Richard, D.; Dumontet, C.; Busserolles, J.; et al. Analgesic and Preventive Effects of Donepezil in Animal Models of Chemotherapy-Induced Peripheral Neuropathy: Involvement of Spinal Muscarinic Acetylcholine M2 Receptors. Biomed. Pharmacother. 2022, 149, 112915. [Google Scholar] [CrossRef]
  34. Zuo, Z.-X.; Wang, Y.-J.; Liu, L.; Wang, Y.; Mei, S.-H.; Feng, Z.-H.; Wang, M.; Li, X.-Y. Huperzine A Alleviates Mechanical Allodynia but Not Spontaneous Pain via Muscarinic Acetylcholine Receptors in Mice. Neural Plast. 2015, 2015, 453170. [Google Scholar] [CrossRef] [PubMed]
  35. Dhanasobhon, D.; Medrano, M.-C.; Becker, L.J.; Moreno-Lopez, Y.; Kavraal, S.; Bichara, C.; Schlichter, R.; Inquimbert, P.; Yalcin, I.; Cordero-Erausquin, M. Enhanced Analgesic Cholinergic Tone in the Spinal Cord in a Mouse Model of Neuropathic Pain. Neurobiol. Dis. 2021, 155, 105363. [Google Scholar] [CrossRef] [PubMed]
  36. Paqueron, X.; Li, X.; Eisenach, J.C. P75-Expressing Elements Are Necessary for Anti-Allodynic Effects of Spinal Clonidine and Neostigmine. Neuroscience 2001, 102, 681–686. [Google Scholar] [CrossRef]
  37. Chen, S.R.; Khan, G.M.; Pan, H.L. Antiallodynic Effect of Intrathecal Neostigmine Is Mediated by Spinal Nitric Oxide in a Rat Model of Diabetic Neuropathic Pain. Anesthesiology 2001, 95, 1007–1012. [Google Scholar] [CrossRef]
  38. Takasu, K.; Honda, M.; Ono, H.; Tanabe, M. Spinal Alpha(2)-Adrenergic and Muscarinic Receptors and the NO Release Cascade Mediate Supraspinally Produced Effectiveness of Gabapentin at Decreasing Mechanical Hypersensitivity in Mice after Partial Nerve Injury. Br. J. Pharmacol. 2006, 148, 233–244. [Google Scholar] [CrossRef] [PubMed]
  39. Marinelli, S.; Vacca, V.; Ricordy, R.; Uggenti, C.; Tata, A.M.; Luvisetto, S.; Pavone, F. The Analgesic Effect on Neuropathic Pain of Retrogradely Transported Botulinum Neurotoxin A Involves Schwann Cells and Astrocytes. PLoS ONE 2012, 7, e47977. [Google Scholar] [CrossRef]
  40. Favre-Guilmard, C.; Auguet, M.; Chabrier, P.-E. Different Antinociceptive Effects of Botulinum Toxin Type A in Inflammatory and Peripheral Polyneuropathic Rat Models. Eur. J. Pharmacol. 2009, 617, 48–53. [Google Scholar] [CrossRef]
  41. Emril, D.R.; Wibowo, S.; Meliala, L.; Susilowati, R. Cytidine 5′-Diphosphocholine Administration Prevents Peripheral Neuropathic Pain after Sciatic Nerve Crush Injury in Rats. J. Pain Res. 2016, 9, 287–291. [Google Scholar] [CrossRef]
  42. Bagdas, D.; Sonat, F.A.; Hamurtekin, E.; Sonal, S.; Gurun, M.S. The Antihyperalgesic Effect of Cytidine-5′-Diphosphate-Choline in Neuropathic and Inflammatory Pain Models. Behav. Pharmacol. 2011, 22, 589–598. [Google Scholar] [CrossRef]
  43. Kanat, O.; Bagdas, D.; Ozboluk, H.Y.; Gurun, M.S. Preclinical Evidence for the Antihyperalgesic Activity of CDP-Choline in Oxaliplatin-Induced Neuropathic Pain. J. BUON 2013, 18, 1012–1018. [Google Scholar]
  44. Kyte, S.L.; Toma, W.; Bagdas, D.; Meade, J.A.; Schurman, L.D.; Lichtman, A.H.; Chen, Z.-J.; Del Fabbro, E.; Fang, X.; Bigbee, J.W.; et al. Nicotine Prevents and Reverses Paclitaxel-Induced Mechanical Allodynia in a Mouse Model of CIPN. J. Pharmacol. Exp. Ther. 2018, 364, 110–119. [Google Scholar] [CrossRef] [PubMed]
  45. Bagdas, D.; Ergun, D.; Jackson, A.; Toma, W.; Schulte, M.K.; Damaj, M.I. Allosteric Modulation of A4β2* Nicotinic Acetylcholine Receptors: Desformylflustrabromine Potentiates Antiallodynic Response of Nicotine in a Mouse Model of Neuropathic Pain. Eur. J. Pain 2018, 22, 84–93. [Google Scholar] [CrossRef]
  46. Abdin, M.J.; Morioka, N.; Morita, K.; Kitayama, T.; Kitayama, S.; Nakashima, T.; Dohi, T. Analgesic Action of Nicotine on Tibial Nerve Transection (TNT)-Induced Mechanical Allodynia through Enhancement of the Glycinergic Inhibitory System in Spinal Cord. Life Sci. 2006, 80, 9–16. [Google Scholar] [CrossRef]
  47. Kiguchi, N.; Kobayashi, Y.; Maeda, T.; Tominaga, S.; Nakamura, J.; Fukazawa, Y.; Ozaki, M.; Kishioka, S. Activation of Nicotinic Acetylcholine Receptors on Bone Marrow-Derived Cells Relieves Neuropathic Pain Accompanied by Peripheral Neuroinflammation. Neurochem. Int. 2012, 61, 1212–1219. [Google Scholar] [CrossRef]
  48. Young, T.; Wittenauer, S.; Parker, R.; Vincler, M. Peripheral Nerve Injury Alters Spinal Nicotinic Acetylcholine Receptor Pharmacology. Eur. J. Pharmacol. 2008, 590, 163–169. [Google Scholar] [CrossRef]
  49. Wieskopf, J.S.; Mathur, J.; Limapichat, W.; Post, M.R.; Al-Qazzaz, M.; Sorge, R.E.; Martin, L.J.; Zaykin, D.V.; Smith, S.B.; Freitas, K.; et al. The Nicotinic A6 Subunit Gene Determines Variability in Chronic Pain Sensitivity via Cross-Inhibition of P2X2/3 Receptors. Sci. Transl. Med. 2015, 7, 287ra72. [Google Scholar] [CrossRef] [PubMed]
  50. Josiah, D.T.; Vincler, M.A. Impact of Chronic Nicotine on the Development and Maintenance of Neuropathic Hypersensitivity in the Rat. Psychopharmacology 2006, 188, 152–161. [Google Scholar] [CrossRef]
  51. Xanthos, D.N.; Beiersdorf, J.W.; Thrun, A.; Ianosi, B.; Orr-Urtreger, A.; Huck, S.; Scholze, P. Role of A5-Containing Nicotinic Receptors in Neuropathic Pain and Response to Nicotine. Neuropharmacology 2015, 95, 37–49. [Google Scholar] [CrossRef]
  52. Brunori, G.; Schoch, J.; Mercatelli, D.; Ozawa, A.; Toll, L.; Cippitelli, A. Influence of Neuropathic Pain on Nicotinic Acetylcholine Receptor Plasticity and Behavioral Responses to Nicotine in Rats. Pain 2018, 159, 2179–2191. [Google Scholar] [CrossRef] [PubMed]
  53. Saika, F.; Kiguchi, N.; Kobayashi, Y.; Kishioka, S. Peripheral Alpha4beta2 Nicotinic Acetylcholine Receptor Signalling Attenuates Tactile Allodynia and Thermal Hyperalgesia after Nerve Injury in Mice. Acta Physiol. (Oxf.) 2015, 213, 462–471. [Google Scholar] [CrossRef]
  54. Holtman, J.R.; Crooks, P.A.; Johnson-Hardy, J.K.; Wala, E.P. The Analgesic and Toxic Effects of Nornicotine Enantiomers Alone and in Interaction with Morphine in Rodent Models of Acute and Persistent Pain. Pharmacol. Biochem. Behav. 2010, 94, 352–362. [Google Scholar] [CrossRef] [PubMed]
  55. Kesingland, A.C.; Gentry, C.T.; Panesar, M.S.; Bowes, M.A.; Vernier, J.M.; Cube, R.; Walker, K.; Urban, L. Analgesic Profile of the Nicotinic Acetylcholine Receptor Agonists, (+)-Epibatidine and ABT-594 in Models of Persistent Inflammatory and Neuropathic Pain. Pain 2000, 86, 113–118. [Google Scholar] [CrossRef] [PubMed]
  56. Kiguchi, N.; Kobayashi, D.; Saika, F.; Matsuzaki, S.; Kishioka, S. Inhibition of Peripheral Macrophages by Nicotinic Acetylcholine Receptor Agonists Suppresses Spinal Microglial Activation and Neuropathic Pain in Mice with Peripheral Nerve Injury. J. Neuroinflamm. 2018, 15, 96. [Google Scholar] [CrossRef] [PubMed]
  57. Ueda, M.; Iida, Y.; Yoneyama, T.; Kawai, T.; Ogawa, M.; Magata, Y.; Saji, H. In Vivo Relationship between Thalamic Nicotinic Acetylcholine Receptor Occupancy Rates and Antiallodynic Effects in a Rat Model of Neuropathic Pain: Persistent Agonist Binding Inhibits the Expression of Antiallodynic Effects. Synapse 2011, 65, 77–83. [Google Scholar] [CrossRef]
  58. Li, W.; Cai, J.; Wang, B.H.; Huang, L.; Fan, J.; Wang, Y. Antinociceptive Effects of Novel Epibatidine Analogs through Activation of A4β2 Nicotinic Receptors. Sci. China Life Sci. 2018, 61, 688–695. [Google Scholar] [CrossRef]
  59. Lynch, J.J.; Wade, C.L.; Mikusa, J.P.; Decker, M.W.; Honore, P. ABT-594 (a Nicotinic Acetylcholine Agonist): Anti-Allodynia in a Rat Chemotherapy-Induced Pain Model. Eur. J. Pharmacol. 2005, 509, 43–48. [Google Scholar] [CrossRef]
  60. Saika, F.; Kiguchi, N.; Matsuzaki, S.; Kobayashi, D.; Kishioka, S. Inflammatory Macrophages in the Sciatic Nerves Facilitate Neuropathic Pain Associated with Type 2 Diabetes Mellitus. J. Pharmacol. Exp. Ther. 2019, 368, 535–544. [Google Scholar] [CrossRef]
  61. Yoon, H.; Kim, M.J.; Yoon, I.; Li, D.X.; Bae, H.; Kim, S.K. Nicotinic Acetylcholine Receptors Mediate the Suppressive Effect of an Injection of Diluted Bee Venom into the GV3 Acupoint on Oxaliplatin-Induced Neuropathic Cold Allodynia in Rats. Biol. Pharm. Bull. 2015, 38, 710–714. [Google Scholar] [CrossRef]
  62. Cheng, L.-Z.; Han, L.; Fan, J.; Huang, L.-T.; Peng, L.-C.; Wang, Y. Enhanced Inhibitory Synaptic Transmission in the Spinal Dorsal Horn Mediates Antinociceptive Effects of TC-2559. Mol. Pain 2011, 7, 56. [Google Scholar] [CrossRef] [PubMed]
  63. Ueda, M.; Iida, Y.; Tominaga, A.; Yoneyama, T.; Ogawa, M.; Magata, Y.; Nishimura, H.; Kuge, Y.; Saji, H. Nicotinic Acetylcholine Receptors Expressed in the Ventralposterolateral Thalamic Nucleus Play an Important Role in Anti-Allodynic Effects. Br. J. Pharmacol. 2010, 159, 1201–1210. [Google Scholar] [CrossRef] [PubMed]
  64. Ji, J.; Bunnelle, W.H.; Anderson, D.J.; Faltynek, C.; Dyhring, T.; Ahring, P.K.; Rueter, L.E.; Curzon, P.; Buckley, M.J.; Marsh, K.C.; et al. A-366833: A Novel Nicotinonitrile-Substituted 3,6-Diazabicyclo[3.2.0]-Heptane Alpha4beta2 Nicotinic Acetylcholine Receptor Selective Agonist: Synthesis, Analgesic Efficacy and Tolerability Profile in Animal Models. Biochem. Pharmacol. 2007, 74, 1253–1262. [Google Scholar] [CrossRef] [PubMed]
  65. Nirogi, R.; Jabaris, S.L.; Jayarajan, P.; Abraham, R.; Shanmuganathan, D.; Rasheed, M.A.; Royapalley, P.K.; Goura, V. Antinociceptive Activity of A4β2* Neuronal Nicotinic Receptor Agonist A-366833 in Experimental Models of Neuropathic and Inflammatory Pain. Eur. J. Pharmacol. 2011, 668, 155–162. [Google Scholar] [CrossRef]
  66. Rueter, L.E.; Kohlhaas, K.L.; Curzon, P.; Surowy, C.S.; Meyer, M.D. Peripheral and Central Sites of Action for A-85380 in the Spinal Nerve Ligation Model of Neuropathic Pain. Pain 2003, 103, 269–276. [Google Scholar] [CrossRef] [PubMed]
  67. Lee, C.-H.; Zhu, C.; Malysz, J.; Campbell, T.; Shaughnessy, T.; Honore, P.; Polakowski, J.; Gopalakrishnan, M. A4β2 Neuronal Nicotinic Receptor Positive Allosteric Modulation: An Approach for Improving the Therapeutic Index of A4β2 nAChR Agonists in Pain. Biochem. Pharmacol. 2011, 82, 959–966. [Google Scholar] [CrossRef] [PubMed]
  68. Gong, S.; Liang, Q.; Zhu, Q.; Ding, D.; Yin, Q.; Tao, J.; Jiang, X. Nicotinic Acetylcholine Receptor A7 Subunit Is Involved in the Cobratoxin-Induced Antinociception in an Animal Model of Neuropathic Pain. Toxicon 2015, 93, 31–36. [Google Scholar] [CrossRef]
  69. Freitas, K.; Ghosh, S.; Ivy Carroll, F.; Lichtman, A.H.; Imad Damaj, M. Effects of A7 Positive Allosteric Modulators in Murine Inflammatory and Chronic Neuropathic Pain Models. Neuropharmacology 2013, 65, 156–164. [Google Scholar] [CrossRef] [PubMed]
  70. Di Cesare Mannelli, L.; Pacini, A.; Matera, C.; Zanardelli, M.; Mello, T.; De Amici, M.; Dallanoce, C.; Ghelardini, C. Involvement of A7 nAChR Subtype in Rat Oxaliplatin-Induced Neuropathy: Effects of Selective Activation. Neuropharmacology 2014, 79, 37–48. [Google Scholar] [CrossRef]
  71. Ji, L.; Chen, Y.; Wei, H.; Feng, H.; Chang, R.; Yu, D.; Wang, X.; Gong, X.; Zhang, M. Activation of Alpha7 Acetylcholine Receptors Reduces Neuropathic Pain by Decreasing Dynorphin A Release from Microglia. Brain Res. 2019, 1715, 57–65. [Google Scholar] [CrossRef]
  72. Bagdas, D.; Wilkerson, J.L.; Kulkarni, A.; Toma, W.; AlSharari, S.; Gul, Z.; Lichtman, A.H.; Papke, R.L.; Thakur, G.A.; Damaj, M.I. The A7 Nicotinic Receptor Dual Allosteric Agonist and Positive Allosteric Modulator GAT107 Reverses Nociception in Mouse Models of Inflammatory and Neuropathic Pain. Br. J. Pharmacol. 2016, 173, 2506–2520. [Google Scholar] [CrossRef]
  73. Arias, H.R.; Ghelardini, C.; Lucarini, E.; Tae, H.-S.; Yousuf, A.; Marcovich, I.; Manetti, D.; Romanelli, M.N.; Elgoyhen, A.B.; Adams, D.J.; et al. (E)-3-Furan-2-Yl-N-p-Tolyl-Acrylamide and Its Derivative DM489 Decrease Neuropathic Pain in Mice Predominantly by A7 Nicotinic Acetylcholine Receptor Potentiation. ACS Chem. Neurosci. 2020, 11, 3603–3614. [Google Scholar] [CrossRef]
  74. Bagdas, D.; Sevdar, G.; Gul, Z.; Younis, R.; Cavun, S.; Tae, H.-S.; Ortells, M.O.; Arias, H.R.; Gurun, M.S. (E)-3-Furan-2-Yl-N-Phenylacrylamide (PAM-4) Decreases Nociception and Emotional Manifestations of Neuropathic Pain in Mice by A7 Nicotinic Acetylcholine Receptor Potentiation. Neurol. Res. 2021, 43, 1056–1068. [Google Scholar] [CrossRef] [PubMed]
  75. Pacini, A.; Di Cesare Mannelli, L.; Bonaccini, L.; Ronzoni, S.; Bartolini, A.; Ghelardini, C. Protective Effect of Alpha7 nAChR: Behavioural and Morphological Features on Neuropathy. Pain 2010, 150, 542–549. [Google Scholar] [CrossRef]
  76. Loram, L.C.; Taylor, F.R.; Strand, K.A.; Maier, S.F.; Speake, J.D.; Jordan, K.G.; James, J.W.; Wene, S.P.; Pritchard, R.C.; Green, H.; et al. Systemic Administration of an Alpha-7 Nicotinic Acetylcholine Agonist Reverses Neuropathic Pain in Male Sprague Dawley Rats. J. Pain 2012, 13, 1162–1171. [Google Scholar] [CrossRef] [PubMed]
  77. Papke, R.L.; Bagdas, D.; Kulkarni, A.R.; Gould, T.; AlSharari, S.D.; Thakur, G.A.; Damaj, M.I. The Analgesic-like Properties of the Alpha7 nAChR Silent Agonist NS6740 Is Associated with Non-Conducting Conformations of the Receptor. Neuropharmacology 2015, 91, 34–42. [Google Scholar] [CrossRef]
  78. Toma, W.; Kyte, S.L.; Bagdas, D.; Jackson, A.; Meade, J.A.; Rahman, F.; Chen, Z.-J.; Del Fabbro, E.; Cantwell, L.; Kulkarni, A.; et al. The A7 Nicotinic Receptor Silent Agonist R-47 Prevents and Reverses Paclitaxel-Induced Peripheral Neuropathy in Mice without Tolerance or Altering Nicotine Reward and Withdrawal. Exp. Neurol. 2019, 320, 113010. [Google Scholar] [CrossRef]
  79. Bagdas, D.; Targowska-Duda, K.M.; López, J.J.; Perez, E.G.; Arias, H.R.; Damaj, M.I. The Antinociceptive and Antiinflammatory Properties of 3-Furan-2-Yl-N-p-Tolyl-Acrylamide, a Positive Allosteric Modulator of A7 Nicotinic Acetylcholine Receptors in Mice. Anesth. Analg. 2015, 121, 1369–1377. [Google Scholar] [CrossRef]
  80. Liu, Z.; Bartels, P.; Sadeghi, M.; Du, T.; Dai, Q.; Zhu, C.; Yu, S.; Wang, S.; Dong, M.; Sun, T.; et al. A Novel α-Conopeptide Eu1.6 Inhibits N-Type (CaV2.2) Calcium Channels and Exhibits Potent Analgesic Activity. Sci. Rep. 2018, 8, 1004. [Google Scholar] [CrossRef] [PubMed]
  81. Wang, D.; Yang, H.; Liang, Y.; Wang, X.; Du, X.; Li, R.; Jiang, Y.; Ye, J. Antinociceptive Effect of Spirocyclopiperazinium Salt Compound DXL-A-24 and the Underlying Mechanism. Neurochem. Res. 2019, 44, 2786–2795. [Google Scholar] [CrossRef] [PubMed]
  82. Napier, I.A.; Klimis, H.; Rycroft, B.K.; Jin, A.H.; Alewood, P.F.; Motin, L.; Adams, D.J.; Christie, M.J. Intrathecal α-Conotoxins Vc1.1, AuIB and MII Acting on Distinct Nicotinic Receptor Subtypes Reverse Signs of Neuropathic Pain. Neuropharmacology 2012, 62, 2202–2207. [Google Scholar] [CrossRef]
  83. Klimis, H.; Adams, D.J.; Callaghan, B.; Nevin, S.; Alewood, P.F.; Vaughan, C.W.; Mozar, C.A.; Christie, M.J. A Novel Mechanism of Inhibition of High-Voltage Activated Calcium Channels by α-Conotoxins Contributes to Relief of Nerve Injury-Induced Neuropathic Pain. Pain 2011, 152, 259–266. [Google Scholar] [CrossRef]
  84. Di Cesare Mannelli, L.; Cinci, L.; Micheli, L.; Zanardelli, M.; Pacini, A.; McIntosh, J.M.; Ghelardini, C. α-Conotoxin RgIA Protects against the Development of Nerve Injury-Induced Chronic Pain and Prevents Both Neuronal and Glial Derangement. Pain 2014, 155, 1986–1995. [Google Scholar] [CrossRef] [PubMed]
  85. Nevin, S.T.; Clark, R.J.; Klimis, H.; Christie, M.J.; Craik, D.J.; Adams, D.J. Are Alpha9alpha10 Nicotinic Acetylcholine Receptors a Pain Target for Alpha-Conotoxins? Mol. Pharmacol. 2007, 72, 1406–1410. [Google Scholar] [CrossRef] [PubMed]
  86. Cai, F.; Xu, N.; Liu, Z.; Ding, R.; Yu, S.; Dong, M.; Wang, S.; Shen, J.; Tae, H.-S.; Adams, D.J.; et al. Targeting of N-Type Calcium Channels via GABAB-Receptor Activation by α-Conotoxin Vc1.1 Variants Displaying Improved Analgesic Activity. J. Med. Chem. 2018, 61, 10198–10205. [Google Scholar] [CrossRef] [PubMed]
  87. Luo, S.; Zhangsun, D.; Harvey, P.J.; Kaas, Q.; Wu, Y.; Zhu, X.; Hu, Y.; Li, X.; Tsetlin, V.I.; Christensen, S.; et al. Cloning, Synthesis, and Characterization of AO-Conotoxin GeXIVA, a Potent A9α10 Nicotinic Acetylcholine Receptor Antagonist. Proc. Natl. Acad. Sci. USA 2015, 112, E4026–E4035. [Google Scholar] [CrossRef]
  88. Wala, E.P.; Crooks, P.A.; McIntosh, J.M.; Holtman, J.R. Novel Small Molecule A9α10 Nicotinic Receptor Antagonist Prevents and Reverses Chemotherapy-Evoked Neuropathic Pain in Rats. Anesth. Analg. 2012, 115, 713–720. [Google Scholar] [CrossRef] [PubMed]
  89. Holtman, J.R.; Dwoskin, L.P.; Dowell, C.; Wala, E.P.; Zhang, Z.; Crooks, P.A.; McIntosh, J.M. The Novel Small Molecule A9α10 Nicotinic Acetylcholine Receptor Antagonist ZZ-204G Is Analgesic. Eur. J. Pharmacol. 2011, 670, 500–508. [Google Scholar] [CrossRef]
  90. Arias, H.R.; Tae, H.-S.; Micheli, L.; Yousuf, A.; Ghelardini, C.; Adams, D.J.; Di Cesare Mannelli, L. Coronaridine Congeners Decrease Neuropathic Pain in Mice and Inhibit A9α10 Nicotinic Acetylcholine Receptors and CaV2.2 Channels. Neuropharmacology 2020, 175, 108194. [Google Scholar] [CrossRef]
  91. Liang, J.; Tae, H.-S.; Zhao, Z.; Li, X.; Zhang, J.; Chen, S.; Jiang, T.; Adams, D.J.; Yu, R. Mechanism of Action and Structure-Activity Relationship of α-Conotoxin Mr1.1 at the Human A9α10 Nicotinic Acetylcholine Receptor. J. Med. Chem. 2022, 65, 16204–16217. [Google Scholar] [CrossRef] [PubMed]
  92. Vincler, M.; Wittenauer, S.; Parker, R.; Ellison, M.; Olivera, B.M.; McIntosh, J.M. Molecular Mechanism for Analgesia Involving Specific Antagonism of Alpha9alpha10 Nicotinic Acetylcholine Receptors. Proc. Natl. Acad. Sci. USA 2006, 103, 17880–17884. [Google Scholar] [CrossRef] [PubMed]
  93. Dyachenko, I.A.; Palikova, Y.A.; Palikov, V.A.; Korolkova, Y.V.; Kazakov, V.A.; Egorova, N.S.; Garifulina, A.I.; Utkin, Y.N.; Tsetlin, V.I.; Kryukova, E.V. α-Conotoxin RgIA and Oligoarginine R8 in the Mice Model Alleviate Long-Term Oxaliplatin Induced Neuropathy. Biochimie 2022, 194, 127–136. [Google Scholar] [CrossRef] [PubMed]
  94. Pacini, A.; Micheli, L.; Maresca, M.; Branca, J.J.V.; McIntosh, J.M.; Ghelardini, C.; Di Cesare Mannelli, L. The A9α10 Nicotinic Receptor Antagonist α-Conotoxin RgIA Prevents Neuropathic Pain Induced by Oxaliplatin Treatment. Exp. Neurol. 2016, 282, 37–48. [Google Scholar] [CrossRef] [PubMed]
  95. Romero, H.K.; Christensen, S.B.; Di Cesare Mannelli, L.; Gajewiak, J.; Ramachandra, R.; Elmslie, K.S.; Vetter, D.E.; Ghelardini, C.; Iadonato, S.P.; Mercado, J.L.; et al. Inhibition of A9α10 Nicotinic Acetylcholine Receptors Prevents Chemotherapy-Induced Neuropathic Pain. Proc. Natl. Acad. Sci. USA 2017, 114, E1825–E1832. [Google Scholar] [CrossRef] [PubMed]
  96. Christensen, S.B.; Hone, A.J.; Roux, I.; Kniazeff, J.; Pin, J.-P.; Upert, G.; Servent, D.; Glowatzki, E.; McIntosh, J.M. RgIA4 Potently Blocks Mouse A9α10 nAChRs and Provides Long Lasting Protection against Oxaliplatin-Induced Cold Allodynia. Front. Cell. Neurosci. 2017, 11, 219. [Google Scholar] [CrossRef] [PubMed]
  97. Huynh, P.N.; Christensen, S.B.; McIntosh, J.M. RgIA4 Prevention of Acute Oxaliplatin-Induced Cold Allodynia Requires A9-Containing Nicotinic Acetylcholine Receptors and CD3+ T-Cells. Cells 2022, 11, 3561. [Google Scholar] [CrossRef]
  98. Huynh, P.N.; Giuvelis, D.; Christensen, S.; Tucker, K.L.; McIntosh, J.M. RgIA4 Accelerates Recovery from Paclitaxel-Induced Neuropathic Pain in Rats. Mar. Drugs 2019, 18, 12. [Google Scholar] [CrossRef]
  99. Gajewiak, J.; Christensen, S.; Dowell, C.; Hararah, F.; Fisher, F.; Huynh, P.N.; Olivera, B.; McIntosh, J.M. Selective Penicillamine Substitution Enables Development of a Potent Analgesic Peptide That Acts Through a Non-Opioid Based Mechanism. J. Med. Chem. 2021, 64, 9271–9278. [Google Scholar] [CrossRef]
  100. Belgi, A.; Burnley, J.V.; MacRaild, C.A.; Chhabra, S.; Elnahriry, K.A.; Robinson, S.D.; Gooding, S.G.; Tae, H.-S.; Bartels, P.; Sadeghi, M.; et al. Alkyne-Bridged α-Conotoxin Vc1.1 Potently Reverses Mechanical Allodynia in Neuropathic Pain Models. J. Med. Chem. 2021, 64, 3222–3233. [Google Scholar] [CrossRef]
  101. Satkunanathan, N.; Livett, B.; Gayler, K.; Sandall, D.; Down, J.; Khalil, Z. Alpha-Conotoxin Vc1.1 Alleviates Neuropathic Pain and Accelerates Functional Recovery of Injured Neurones. Brain Res. 2005, 1059, 149–158. [Google Scholar] [CrossRef]
  102. Wang, H.; Li, X.; Zhangsun, D.; Yu, G.; Su, R.; Luo, S. The A9α10 Nicotinic Acetylcholine Receptor Antagonist AO-Conotoxin GeXIVA[1,2] Alleviates and Reverses Chemotherapy-Induced Neuropathic Pain. Mar. Drugs 2019, 17, 265. [Google Scholar] [CrossRef]
  103. Li, Z.; Han, X.; Hong, X.; Li, X.; Gao, J.; Zhang, H.; Zheng, A. Lyophilization Serves as an Effective Strategy for Drug Development of the A9α10 Nicotinic Acetylcholine Receptor Antagonist α-Conotoxin GeXIVA[1,2]. Mar. Drugs 2021, 19, 121. [Google Scholar] [CrossRef]
  104. Li, X.; Hu, Y.; Wu, Y.; Huang, Y.; Yu, S.; Ding, Q.; Zhangsun, D.; Luo, S. Anti-Hypersensitive Effect of Intramuscular Administration of AO-Conotoxin GeXIVA[1,2] and GeXIVA[1,4] in Rats of Neuropathic Pain. Prog. Neuropsychopharmacol. Biol. Psychiatry 2016, 66, 112–119. [Google Scholar] [CrossRef] [PubMed]
  105. Song, Z.; Meyerson, B.A.; Linderoth, B. Muscarinic Receptor Activation Potentiates the Effect of Spinal Cord Stimulation on Pain-Related Behavior in Rats with Mononeuropathy. Neurosci. Lett. 2008, 436, 7–12. [Google Scholar] [CrossRef] [PubMed]
  106. Wang, Y.-J.; Zuo, Z.-X.; Zhang, M.; Feng, Z.-H.; Yan, M.; Li, X.-Y. The Analgesic Effects of (5R,6R)6-(3-Propylthio-1,2,5-Thiadiazol-4-Yl)-1-Azabicyclo[3.2.1] Octane on a Mouse Model of Neuropathic Pain. Anesth. Analg. 2017, 124, 1330–1338. [Google Scholar] [CrossRef] [PubMed]
  107. Ortega-Legaspi, J.M.; López-Avila, A.; Coffeen, U.; del Angel, R.; Pellicer, F. Scopolamine into the Anterior Cingulate Cortex Diminishes Nociception in a Neuropathic Pain Model in the Rat: An Interruption of “Nociception-Related Memory Acquisition”? Eur. J. Pain 2003, 7, 425–429. [Google Scholar] [CrossRef]
  108. Koga, K.; Matsuzaki, Y.; Migita, K.; Shimoyama, S.; Eto, F.; Nakagawa, T.; Matsumoto, T.; Terada, K.; Mishima, K.; Furue, H.; et al. Stimulating Muscarinic M1 Receptors in the Anterior Cingulate Cortex Reduces Mechanical Hypersensitivity via GABAergic Transmission in Nerve Injury Rats. Brain Res. 2019, 1704, 187–195. [Google Scholar] [CrossRef] [PubMed]
  109. Wood, M.W.; Martino, G.; Coupal, M.; Lindberg, M.; Schroeder, P.; Santhakumar, V.; Valiquette, M.; Sandin, J.; Widzowski, D.; Laird, J. Broad Analgesic Activity of a Novel, Selective M1 Agonist. Neuropharmacology 2017, 123, 233–241. [Google Scholar] [CrossRef]
  110. Calcutt, N.A.; Smith, D.R.; Frizzi, K.; Sabbir, M.G.; Chowdhury, S.K.R.; Mixcoatl-Zecuatl, T.; Saleh, A.; Muttalib, N.; Van der Ploeg, R.; Ochoa, J.; et al. Selective Antagonism of Muscarinic Receptors Is Neuroprotective in Peripheral Neuropathy. J. Clin. Investig. 2017, 127, 608–622. [Google Scholar] [CrossRef]
  111. Ghazisaeidi, S.; Muley, M.M.; Salter, M.W. Neuropathic Pain: Mechanisms, Sex Differences, and Potential Therapies for a Global Problem. Annu. Rev. Pharmacol. Toxicol. 2023, 63, 565–583. [Google Scholar] [CrossRef]
  112. Mapplebeck, J.C.S.; Beggs, S.; Salter, M.W. Sex Differences in Pain: A Tale of Two Immune Cells. Pain 2016, 157, S2–S6. [Google Scholar] [CrossRef]
  113. Salehi, B.; Sestito, S.; Rapposelli, S.; Peron, G.; Calina, D.; Sharifi-Rad, M.; Sharopov, F.; Martins, N.; Sharifi-Rad, J. Epibatidine: A Promising Natural Alkaloid in Health. Biomolecules 2018, 9, 6. [Google Scholar] [CrossRef]
  114. Kennedy, A.C.; Belgi, A.; Husselbee, B.W.; Spanswick, D.; Norton, R.S.; Robinson, A.J. α-Conotoxin Peptidomimetics: Probing the Minimal Binding Motif for Effective Analgesia. Toxins 2020, 12, 505. [Google Scholar] [CrossRef] [PubMed]
  115. Wilkerson, J.L.; Deba, F.; Crowley, M.L.; Hamouda, A.K.; McMahon, L.R. Advances in the In Vitro and In Vivo Pharmacology of Alpha4beta2 Nicotinic Receptor Positive Allosteric Modulators. Neuropharmacology 2020, 168, 108008. [Google Scholar] [CrossRef] [PubMed]
  116. Bagdas, D.; Gurun, M.S.; Flood, P.; Papke, R.L.; Damaj, M.I. New Insights on Neuronal Nicotinic Acetylcholine Receptors as Targets for Pain and Inflammation: A Focus on A7 nAChRs. Curr. Neuropharmacol. 2018, 16, 415–425. [Google Scholar] [CrossRef] [PubMed]
  117. Shiers, S.; Klein, R.M.; Price, T.J. Quantitative Differences in Neuronal Subpopulations between Mouse and Human Dorsal Root Ganglia Demonstrated with RNAscope in Situ Hybridization. Pain 2020, 161, 2410–2424. [Google Scholar] [CrossRef] [PubMed]
  118. Umana, I.C.; Daniele, C.A.; McGehee, D.S. Neuronal Nicotinic Receptors as Analgesic Targets: It’s a Winding Road. Biochem. Pharmacol. 2013, 86, 1208–1214. [Google Scholar] [CrossRef]
  119. Gotti, C.; Clementi, F. Neuronal Nicotinic Receptors: From Structure to Pathology. Prog. Neurobiol. 2004, 74, 363–396. [Google Scholar] [CrossRef]
  120. Del Bufalo, A.; Cesario, A.; Salinaro, G.; Fini, M.; Russo, P. Alpha9 Alpha10 Nicotinic Acetylcholine Receptors as Target for the Treatment of Chronic Pain. Curr. Pharm. Des. 2014, 20, 6042–6047. [Google Scholar] [CrossRef]
  121. Ditre, J.W.; Heckman, B.W.; Zale, E.L.; Kosiba, J.D.; Maisto, S.A. Acute Analgesic Effects of Nicotine and Tobacco in Humans: A Meta-Analysis. Pain 2016, 157, 1373–1381. [Google Scholar] [CrossRef]
  122. Hahn, E.J.; Rayens, M.K.; Kirsh, K.L.; Passik, S.D. Brief Report: Pain and Readiness to Quit Smoking Cigarettes. Nicotine Tob. Res. 2006, 8, 473–480. [Google Scholar] [CrossRef]
  123. Pandya, A.A.; Yakel, J.L. Effects of Neuronal Nicotinic Acetylcholine Receptor Allosteric Modulators in Animal Behavior Studies. Biochem. Pharmacol. 2013, 86, 1054–1062. [Google Scholar] [CrossRef] [PubMed]
  124. Tan, C.-C.; Yu, J.-T.; Wang, H.-F.; Tan, M.-S.; Meng, X.-F.; Wang, C.; Jiang, T.; Zhu, X.-C.; Tan, L. Efficacy and Safety of Donepezil, Galantamine, Rivastigmine, and Memantine for the Treatment of Alzheimer’s Disease: A Systematic Review and Meta-Analysis. J. Alzheimer’s Dis. 2014, 41, 615–631. [Google Scholar] [CrossRef] [PubMed]
  125. Hartmann-Boyce, J.; Chepkin, S.C.; Ye, W.; Bullen, C.; Lancaster, T. Nicotine Replacement Therapy versus Control for Smoking Cessation. Cochrane Database Syst. Rev. 2018, 2018, CD000146. [Google Scholar] [CrossRef] [PubMed]
  126. Tonstad, S.; Arons, C.; Rollema, H.; Berlin, I.; Hajek, P.; Fagerström, K.; Els, C.; McRae, T.; Russ, C. Varenicline: Mode of Action, Efficacy, Safety and Accumulated Experience Salient for Clinical Populations. Curr. Med. Res. Opin. 2020, 36, 713–730. [Google Scholar] [CrossRef]
  127. Nair, A. Publication Bias—Importance of Studies with Negative Results! Indian J. Anaesth. 2019, 63, 505. [Google Scholar] [CrossRef]
Figure 1. PRISMA flow chart of study selection.
Figure 1. PRISMA flow chart of study selection.
Pharmaceuticals 16 01363 g001
Figure 2. Pharmacological targets identified in the selected publications (image: Wikipedia).
Figure 2. Pharmacological targets identified in the selected publications (image: Wikipedia).
Pharmaceuticals 16 01363 g002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Montigné, E.; Balayssac, D. Exploring Cholinergic Compounds for Peripheral Neuropathic Pain Management: A Comprehensive Scoping Review of Rodent Model Studies. Pharmaceuticals 2023, 16, 1363. https://doi.org/10.3390/ph16101363

AMA Style

Montigné E, Balayssac D. Exploring Cholinergic Compounds for Peripheral Neuropathic Pain Management: A Comprehensive Scoping Review of Rodent Model Studies. Pharmaceuticals. 2023; 16(10):1363. https://doi.org/10.3390/ph16101363

Chicago/Turabian Style

Montigné, Edouard, and David Balayssac. 2023. "Exploring Cholinergic Compounds for Peripheral Neuropathic Pain Management: A Comprehensive Scoping Review of Rodent Model Studies" Pharmaceuticals 16, no. 10: 1363. https://doi.org/10.3390/ph16101363

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop