Next Article in Journal
Phloridzin Reveals New Treatment Strategies for Liver Fibrosis
Next Article in Special Issue
Difelikefalin in the Treatment of Chronic Kidney Disease-Associated Pruritus: A Systematic Review
Previous Article in Journal
Analgesic and Anesthetic Efficacy of Rocuronium/Sugammadex in Otorhinolaryngologic Surgery: A Propensity Score-Matched Analysis
Previous Article in Special Issue
Chronic Pruritus in Atopic Patients Treated with Dupilumab: Real Life Response and Related Parameters in 354 Patients
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Druggable Targets and Compounds with Both Antinociceptive and Antipruritic Effects

1
Department of Dermatology, Taipei Medical University-Shuang Ho Hospital, New Taipei City 23561, Taiwan
2
Department of Dermatology, School of Medicine, College of Medicine, Taipei Medical University, Taipei 11031, Taiwan
3
International Ph.D. Program for Cell Therapy and Regeneration Medicine, College of Medicine, Taipei Medical University, Taipei 11031, Taiwan
4
Department of Dermatology, Faculty of Medicine, University of Medicine and Pharmacy at Ho Chi Minh City, Ho Chi Minh City 70000, Vietnam
5
Department of Dermatology, National Taiwan University Hospital, Taipei 100225, Taiwan
*
Author to whom correspondence should be addressed.
Pharmaceuticals 2022, 15(7), 892; https://doi.org/10.3390/ph15070892
Submission received: 3 June 2022 / Revised: 7 July 2022 / Accepted: 15 July 2022 / Published: 19 July 2022

Abstract

:
Pain and itch are both important manifestations of various disorders, such as herpes zoster, atopic dermatitis, and psoriasis. Growing evidence suggests that both sensations have shared mediators, overlapping neural circuitry, and similarities in sensitization processes. In fact, pain and itch coexist in some disorders. Determining pharmaceutical agents and targets for treating pain and itch concurrently is of scientific and clinical relevance. Here we review the neurobiology of pain and itch and discuss the pharmaceutical targets as well as novel compounds effective for the concurrent treatment of these sensations.

1. Introduction

Pain and itch are regarded as two sides of the same coin; painful stimuli normally diminish itch, as it is commonly perceived that “painful” scratching relieves the sensation of itch. Several distinct and specific pathways for pain and itch processing have been discovered, and multiple molecules have been identified in specific neurons involved in the processing of these sensations. Growing evidence indicates that pain and itch have common pathways and properties, including shared mediators and overlapping circuitry in primary afferents. In particular, great overlap in the activity and sensitization of chronic pain and itch is observable. The coexistence of chronic pain and itch is clinically common, such as in psoriasis and atopic dermatitis, two of the most prevalent chronic inflammatory skin diseases, as well as in postherpetic neuralgia [1,2,3]. Some investigations on patients with atopic dermatitis have reported that subjects experiencing concomitant pain and itch accounted for about 59–78% of total cases [4,5]. For plaque psoriasis, one study by Honma et al. showed that 60.3% of patients reported both pain and itch on the skin [1]. A study to measure the impact level of pain and itch on the daily tasks of patients with herpes zoster over 50 years old observed that the proportion of patients having both pain and itch was 67% at inclusion time [6]. Moreover, the percentage of patients suffering from postherpetic neuralgia associated with itch was in the range of 30–58% [3]. Indeed, chronic pain and itch are the most common disorders in clinical practice, and treatments are generally disappointing and unsuccessful [7,8]. Furthermore, persistent pain or itch can generate corresponding disturbances in neural circuitry that pathologically amplify the problem, substantially affecting the quality of life [9]. These matters indicate that common characteristics and pathways in processing itch and pain could be implicated and become a barrier in managing their chronic condition. Therefore, this review will highlight the commonality of two aversive sensations and recent advances in treatments targeting both sensations.

2. Overlapping Nature of Pain and Itch

Pain and itch share many overlapping properties in various aspects. In terms of evolutionary significance, both pain and itch sensations are hypothesized to be projective warnings against potential harm by inducing appropriate reflexes and generating aversive responses. Anatomically, both pain and itch are transmitted via the C fibers of primary sensory afferents. Clinically, patients with chronic pain and itch often present with multiple similar manifestations. For example, some patients report attentional focus on the body-related stimuli, which leads to hypervigilance and hypersensitivity via peripheral and central sensitization [10,11,12,13,14,15]. In sum, chronic pain and itch have various similarities. Pruritogens may activate nociceptive primary afferent fibers and generate simultaneous pruritic and nociceptive sensations, as well as body sensations leading to hypersensitivity. They can also induce central and peripheral sensitization [16]. Furthermore, surgical removal of the ventral lateral funiculus of the spinal cord has been demonstrated to attenuate itch and pain concomitantly, and some reports have observed that itch cannot be elicited in patients with congenital insensitivity to pain [17,18,19]. In the past, some cancer patients suffering from intractable pain underwent anterior cordotomy to interrupt the transduction pathway of pain. This classical method resulted in some significant improvement and enhancement of pain thresholds, especially for sharp pain [20,21]. After that, Hyndman et al. observed that cordotomy patients also reported reduced sensation to pruritogens [22]. Interestingly, the attenuated response to itch only occurred in the analgesic regions, and this sensation was produced normally in regions with normal pain [22]. To sufficiently capture the complexity of the pathogenesis of pain and itch, understanding their mechanisms and commonalities is essential. Thus, novel strategies can be developed to target the refractory chronicity of pain and itch.

2.1. Overlapping of Algogens and Pruritogens

Multiple peripheral and spinal neurons are involved in the pathogenesis of both pain and itch [18,23,24,25,26,27,28,29]. In fact, some pruritogens can evoke sensations and behaviors associated with both pain and itch [30]. For instance, histamine and capsaicin are at times considered “partial pruritogens”, as these agents can induce a combination of behaviors including both scratching and wiping, depending on how and where they are applied [31,32]. When histamine is applied on the skin, the sensation of itch is perceived. However, when histamine is injected into the underlying tissue, pain is perceived instead. Notably, a psychophysical study by Sikand et al. demonstrated the coexistence of pain and itch when canonical “pruritogens” or “algogens” were administered into human skin [16]. In this study, the intradermal injection of histamine, a canonical pruritogen, induced itch along with transient pain sensations such as pricking and burning. Conversely, the intradermal injection of capsaicin, a canonical algogen, evoked itch along with a pain sensation in a concentration-dependent manner [16]. Other pruritogens may also excite or sensitize a subset of nociceptors, including SLIGRL (a peptide derived from the N-terminus of protease-activated receptor-2), serotonin, acetylcholine, endothelin-1, chloroquine, BAM8-22, prostaglandin-E2, and bradykinin [33]. This evidence demonstrates the overlapping nature of itch and pain and suggests therapeutic targets for the treatment of pain and itch at the same time.

2.2. Central and Peripheral Sensitization in Pathological Itch and Pain Conditions

Under normal physiological conditions, the sensations of pain and itch are considerably distinct (Figure 1). However, under pathological conditions, the chronicity of these sensations blurs the differentiation of these sensations, complicating their treatments (Figure 2). In fact, painful stimuli may be no longer sufficiently strong to inhibit itch and may even potentiate itch in conditions characterized by chronic pruritus because of neural plasticity. Moreover, magnetic resonance imaging study demonstrated that perception of pain and itch is governed by overlapping areas of the brain [34].
Chronic pain and chronic itch involve a similar disruption in the peripheral sensitization of the primary afferent neurons and central sensitization of the central nervous system [35]. This leads to hypersensitivity or the inappropriate hyperactivation of nociceptive and pruriceptive neurons [36,37,38,39,40]. Persistent pain and itch, which are both aversive sensations, have several correlated clinical manifestations. If uncontrolled, they exert a strong negative impact on the patient’s quality of life. Notably, several mediators are engaged in both chronic itch and chronic pain, including opioids, proteases, substance P, nerve growth factor (NGF), neurotrophin 4, and their respective receptors, including µ-opioid receptor (MOR), k-opioid receptors (KOR), protease-activated receptor 2 (PAR-2), tyrosine kinase receptor (TrkA), transient receptor potential (TRP) channels, and cannabinoid receptors [34]. Understanding the similarities between central and peripheral sensitization in chronic itch and chronic pain may aid in the development of effective treatments.

2.2.1. Peripheral Sensitization

The sensitization of the peripheral nervous system may also contribute to chronic itch [41]. The abnormal activation of primary sensory afferents can be partly explained by certain aberrant characteristics, including the hyperinnervation of intraepidermal skin fibers and lower tolerance thresholds in chronic itch and chronic pain [35,42,43,44,45,46]. In the context of chronic itch, common pain-evoking triggers, such as electrical, mechanical, chemical, and heat activators, provoke itch instead of pain in the affected skin regions. Similarly, the classic inflammatory algogens, including bradykinin, serotonin, tumor necrosis factor-alpha (TNF-α), and prostaglandins, may also activate or sensitize pruriceptors and provoke itch [47,48,49].
Notably, elevated nerve growth factor (NGF) levels in the skin and blood have been reported in patients with atopic dermatitis, contact dermatitis, and chronic prurigo, as well as in the pruritic lesions of the patients with psoriasis [50,51,52,53]. Elevated NGF levels have also been observed in painful disorders, including chronic localized pain and vulvar dysesthesia. This finding is consistent with observations that NGF may stimulate nociceptors [34,54]. In addition, NGF may activate peripheral histamine-sensitive C-fibers. These fibers are engaged in the signaling of itch, and the serum levels of NGF are correlated with the clinical severity of atopic dermatitis [50]. Taken together, theoretically anti-NGF therapy might become the treatment target for refractory chronic pain and chronic itch despite its failures in clinical trials on osteoarthritic pain [55,56]. The similarities between localized pain and pruritus reveal the analogous pathogenesis of the activity and the sensitization of nerve endings in peripheral disorders characterized by these two sensations.

2.2.2. Central Sensitization

Chronic pain and chronic itch have similar central sensitization mechanisms. These conditions can result from poor inhibitory control and increased excitatory synaptic transmission. Changes in central sensitization may lead to abnormal sensations, such as allokinesis and hyperkinesis in chronic itch, and allodynia and hyperalgesia in chronic pain. These sensory abnormalities involve the body being sensitized to innocuous mechanical stimuli that do not normally provoke pain and itch [44,57,58,59].
There are parallel phenomena for chronic pain and chronic itch via neuroplasticity changes in the spinal cord. Specifically, allokinesis in chronic itch can occur due to light-touch-elicited itch via Aβ-LTMRs (low-threshold mechanoreceptors) in the adjacent unaffected skin. The maintenance of this activation requires the continual activity of primary C-fibers and nociceptors [60,61]. This phenomenon corresponds to the allodynia that affects the uninjured surroundings of nerve injuries in chronic pain. Another manifestation of hypersensitivity in persistent itch is hyperkinesis, in which a punctate mechanical stimulus, such as a pinprick, triggers an intense itch sensation in regions in the vicinity of the affected itchy skin. Hyperknesis is similar to hyperalgesia, a chronic pain phenomenon in which slightly painful stimulation of the sites adjacent to a lesion induces extreme pain [57,62,63]. These phenomena of hypersensitivity can be partly explained by neural plasticity within the dorsal spinal cord leading to central sensitization via C-fibers and Aβ fibers [64,65,66,67].
Pruritogens are recognized as painful stimuli in some chronic neuropathic pain disorders. Supporting these observations, the application of histamine iontophoresis was perceived as burning pain instead of itch in patients suffering from neuropathic pain. It suggested an abnormal spinal hypersensitivity to C-fiber input in chronic pain [68,69]. In addition, in the context of the neuropathic itch of brachioradial pruritus, some patients reported symptoms resembling neuropathic pain, including burning, stinging, and tingling [70]. Conversely, patients with chronic itch demonstrated an itch reflex to canonical painful stimuli, such as repetitive electrical shock, bradykinin, and acetylcholine [24,44,71]. This mechanism leads to a vicious itch-scratch cycle in patients with atopic dermatitis [72]. Furthermore, painful electrical stimuli, acetylcholine, and bradykinin may provoke itch instead of pain in patients with atopic dermatitis [71,73]. Interestingly, a significant proportion of patients complain about postherpetic itch instead of pain, and more than 25% of patients suffering from postherpetic pain also report itch [3,74,75]. In this respect, medicines for neuropathic pain, such as gabapentin, were also reported to attenuate neuropathic itch [76].

3. Therapeutic Targets for Pain and Itch

Considering the overlapping mechanisms of pain and itch, multiple therapeutic targets and their derived compounds exist with both antipruritic and analgesic effects. The existence of such compounds reflects the complex nature of the association in the pathogenesis of pain and itch. Here we enlist the druggable targets for therapeutics effective in both pain and itch, with reviews for novel and relevant compounds in the preclinical testing of these therapeutics (Table 1).

3.1. TRPV1 (Transient Receptor Potential Vanilloid 1)

TRPV1 was initially identified as a marker for pain-sensing neurons and remains one of the most investigated channels in the thermoTRP family [77]. It is known to be activated by capsaicin and noxious heat with subsequent neuroinflammatory processes [78]. Mice deficient in TRPV1 demonstrate a drastic reduction in various pain-associated behaviors [79]. Moreover, evidence from rodents and monkeys indicates that TRPV1 is also responsible for histamine-dependent itch [80,81].
Several TRPV1 antagonists have been shown to alleviate both pain and itch. Due to severe side effects from the systemic administration of TRPV1 antagonists, currently, the strategy of TRPV1 channel blockade has been focused on topical applications. One example is asivatrep (PAC-14028), a TRPV1 antagonist, reduces both pain and itch behaviors in animal models. It has been demonstrated that asivatrep reduces pain in inflammatory bowel disease, diabetic neuropathy, and visceral pain [82]. In a mouse model of atopic dermatitis, asivatrep alleviated itch along with decreased mast cell degranulation and neuroinflammation [83]. In a Phase III trial, the topical application of asivatrep reduced itch and skin inflammation in adult patients with atopic dermatitis [84]. Similarly, another TRPV1 antagonist with modified “soft drug” properties, AG1529, demonstrated efficacy in ameliorating histamine-induced pain and itch in animal models [85].
Interestingly, topical capsaicin may also alleviate pain and itch as a TRPV1 agonist, with a proposed mechanism involving the extended desensitization of sensory neurons. For example, NGX-4010, the active ingredient of 8% trans-capsaicin patches, is effective in treating pain in neuropathies [86,87]. Previous studies have also shown that 0.025% and 0.075% capsaicin creams may be effective in the treatment of arthritic pain [87]. Evidence suggests that topical capsaicin is also effective in the treatment of itch in various dermatoses, such as postherpetic pruritus, uremia, notalgia paresthetica, and chronic prurigo [88,89].

3.2. TRPV3 (Transient Receptor Potential Vanilloid 3)

TRPV3 is a temperature-sensitive channel expressed in the neuronal and non-neuronal structures of the skin, including sensory neurons, keratinocytes, hair shafts, and vasculature [90]. Although the exact roles of TRPV3 in pain and itch remain to be elucidated, several lines of evidence indicate that TRPV3 is involved at least in the modulation of pain and itch, as well as in the development of dermatitis. Mice deficient in TRPV3 displayed reduced pain response to noxious heat [91]. Transgenic mice carrying a point mutation, Gly573Ser, leading to overexpression in TRPV3 showed increased itch behaviors and peripheral neurite outgrowth in the skin [92]. Therefore, TRPV3 may be an ideal pharmacological target for the concurrent inhibition of pain and itch. However, drug development targeting TRPV3 is sluggish due to a lack of successful compounds progressing into clinical trials. This paper reviews several novel compounds with preclinical data suggesting promise in the concurrent targeting of pain and itch.
One example is citrusinine-II, an acridone alkaloid derived from the plant Atalantia monophyla [93]. It is a TRPV3 inhibitor that targets its S4 helix. In the mouse model, citrusinine-II demonstrates antipruritic activities against itch induced by histamine and dry skin, but not acute itch induced by chloroquine. Moreover, citrusinine-II possesses antinociceptive properties against pain evoked by noxious heat and acetic acid [93].
Dyclonine, an over-the-counter oral anesthetic for pain from sore throats, exerts its antinociceptive and antipruritic effects through TRPV3 inhibition [94,95]. Electrophysiological studies have implicated dyclonine as a potent TRPV3 inhibitor under the thermal or chemical activation of TRPV3. In mouse models, dyclonine administration reduced TRPV3-associated pain and itch behaviors. Hydra Biosciences’ FTP-THQ, a selective TRPV3 antagonist, also showed favorable efficacy in relieving histamine-induced itch and formalin-induced pain in rodents [96,97].

3.3. TRPV4 (Transient Receptor Potential Vanilloid 4)

TRPV4 is widely expressed in the body, including the sensory neurons, immune cells, skin, vasculature, etc. Notably, it serves as a sensor for osmolarity, as well as chemical, thermal, and mechanical triggers [98]. TRPV4 is known to be involved in pain and itch. Mice deficient in TRPV4 exhibited attenuated pain responses to hypotonicity and formalin [99,100]. They also exhibited attenuated itch responses to chloroquine, histamine, and serotonin [101,102].
TRPV4 blockade has attracted interest for its potential in treating both pain and itch. Isopropyl cyclohexane, along with its derivatives, is a novel compound. They were patented by the National Institute of Natural Science in Japan for their efficacy in preventing pain and itch in hypersensitive skin [103]. A series of compounds have also been patented by the Liedtke’s group to manage pain and itch [103].

3.4. TRPA1 (Transient Receptor Potential Cation Channel Subfamily A Member 1)

Since its discovery in 2003, TRPA1 has garnered substantial attention for its potential in pharmaceutical development [104]. Expressed as a subset of TRPV1-expressing sensory neurons, TRPA1 is responsible for chemically, thermally, and mechanically evoked pain and itch. It is also expressed in the non-neuronal tissues such as fibroblasts, pancreas, lung, urinary tract, and endothelium. In several animal studies, TRPA1-deficient mice exhibited attenuated pain responses to formalin, noxious cold, and tactile force [105,106,107,108]. Corresponding evidence was reported in a study of gain-of-function N855S mutation on TRPA1 in patients with familial episodic pain syndrome [109]. In a mouse model, TRPA1 was found responsible for histamine-independent itch by Mas-related G-protein coupled receptor member A3 (MrgprA3) as a downstream effector [108]. A psychophysical study in human observed that a moderate itch sensation was evoked by the topical application of trans-cinnamaldehyde, a TRPA1 agonist [110].
Unlike TRPV1 antagonists, the systemic administration of TRPA1 antagonists does not induce major adverse effects. However, currently, all known clinical trials on TRPA1 antagonists have been discontinued, possibly due to pharmacokinetic issues and poor mouse-to-human translational results from intrinsic species-dependent properties in TRPA1 per se [111]. One potential novel approach for treating pain and itch via TRPA1-mediated mechanisms involves the manipulation of the TRPA1-V1 complex. Tmem100, a regulator of the TRPA1-V1 complex, is associated with both TRPA1-related pain behaviors and itch and was patented for their management [112,113].

3.5. TRPM8 (Transient Receptor Potential Cation Channel Subfamily M Member 8)

TRPM8 is a cold-sensing channel in a subset of sensory neurons [114,115]. Several natural and synthetic cooling agents have been identified as its agonists, such as menthol, icilin, rotundifolone, eucalyptol, borneol, etc. [116]. These canonical TRPM8 agonists have been shown to alleviate both pain and itch. Mounting evidence from both human and animal studies indicates that the application of natural TRPM8 agonists such as menthol, eucalyptol, and icilin may attenuate the chronic pain of widely different etiologies. The analgesic effect of TRPM8 agonists is likely due to the desensitization of nociceptive neurons following initial activation. These include inflammatory pain, migraine, and neuropathic pain from chemotherapy, nerve injury, and herpes zoster [117,118,119,120,121,122]. Meanwhile, menthol and peppermint oil have traditionally been employed as antipruritics. A study by Palkar et al. unveiled the neural circuitry for TRPM8-mediated antipruritic effects and provided a solid basis for its mechanism [123].
Aside from canonical TRPM8 agonists, several novel compounds with TRPM8 agonist properties demonstrate great potential for treating both pain and itch. For example, di-isopropyl-phosphinoyl-alkane (DAPA), a synthetic TRPM8 agonist, has been patented for treating sensory dysfunction, including pain and itch [124]. Another synthetic TRPM8 agonist, WS-12, exhibited analgesic effects in animal models and also possesses the potential for treating pruritus [125].

3.6. TRPC3 (Transient Receptor Potential Channel Subfamily C Member 3)

Emerging evidence has revealed that TRPC3 is a potential target for both pain and itch treatment. It is expressed ubiquitously; in particular, in the somatosensory system, TRPC3 is expressed in a subset of non-peptidergic neurons. Pharmacological and genetic studies in murine models suggest that TRPC3 is involved in pain and nonhistaminergic itch [126]. Hence, TRPC3 may be another candidate for the development of antinociceptives and antipruritics in the future.

3.7. Kappa and Miu Opioid Receptors

Opioid receptors consist of a family of G protein-coupled receptors with various physiological functions. They are widely distributed in the central and peripheral systems, as well as the digestive tract. There are four types of opioid receptors, including μ-opioid receptor (MOR), k-opioid receptor (KOR), δ-opioid receptor (DOR), and opioid receptor-like 1 receptor. Among them, KOR and MOR are known to be involved in sensations of pain and itch. Intrathecal administration of morphine induces itch in 10–20% of obstetric patients [127]. One classic pharmacology study with various opioid receptor blockers by Ko et al. demonstrated that opioid-induced itch is mediated by MOR in the central nervous system [128]. In the skin and associated peripheral nervous system, an imbalance of KOR and MOR has been postulated to signal itch, as the agonists of MOR provoke itch, and the agonists of KOR inhibit itch [129].
The analgesic effects of opioids, including morphine, oxycodone, oxymorphone, and fentanyl, are mediated by MOR. However, opioids also bring about major adverse effects, including addiction and respiratory depression. Given that KOR is not associated with such adverse effects, much effort has been made into the development of selective and peripherally acting KOR agonists for the treatment of pain and itch. For example, difelikefalin (CR845), a peptide-derived, peripherally acting KOR agonist, demonstrates great efficacy for uremic pruritus and postsurgical pain [130,131]. Another novel KOR agonist, HSK21542, has shown promising results for antinociceptive and antipruritic effects in animal models [132]. Similarly, triazole 1.1, a synthetic KOR agonist, displayed both antinociceptive and antipruritic effects in rodents [133].
Several canonical opioids also demonstrate efficacy in the treatment of both pain and itch. Naloxone and naltrexone, the opioid receptor antagonists with the greatest affinity against MOR, have been reported to be effective in the treatment of itch in various diseases [134,135]. The administration of these agents alone has been found to enhance pain [136]. However, when these agents are administered with buprenorphine, a partial MOR agonist, they have treatment effects on pain, particularly in patients with opioid addiction [137]. Another example is butorphanol, a partial agonist and antagonist for MOR and an agonist for KOR. Butorphanol is widely used as an analgesic. Studies have noted that butorphanol may be effective for the treatment of itch [138,139].
Nalbuphine is another promising candidate for the treatment of pain and itch at the same time. As a KOR agonist and partial MOR agonist/antagonist, nalbuphine possesses both antinociceptive and antipruritic properties. It is commercially available for managing pain. Additionally, it has been shown to be effective in uremic pruritus, opioid-induced pruritus, and chronic prurigo in various studies [140,141,142].

3.8. Histamine Receptors

Histamine receptors belong to the G protein-coupled receptor family and consist of four types of receptors, H1R, H2R, H3R, and H4R. These receptors are widely expressed in the nervous system, immune cells, fibroblasts, endothelium, epithelium, etc. [143]. The most important physiological functions for histamine receptors are the modulation of inflammation, as well as the signaling of itch circuitry. Histamine has long been known to induce the triple response of Lewis, which is characterized by wheals and inflammatory changes, and intradermal injection of histamine evokes the sensation of itch [144,145]. The prescription of antihistamines has become a standard and traditional treatment for itch, although the effects are at times suboptimal. Among the four receptors, H1R and H4R are most involved in both pain and itch.

3.8.1. Histamine H1 Receptor (H1R)

H1R is an excitatory receptor expressed in sensory neurons and responsible for histamine-dependent itch via mechanical insensitive fibers (CMIA) with TRPV1 as a downstream effector [80]. Although not clinically or widely recognized as a common target for analgesics, H1R has long been implicated in the pathogenesis of neuropathic pain in various animal models. Antihistamines targeting H1R, including chlorpheniramine, fexofenadine, and promethazine, have shown efficacy in attenuating pain in different neuropathic pain models in rodents [146]. Several H1R antagonists, such as diphenhydramine, orphenadrine, mepyramine, and pyrilamine, have been reported to be clinically beneficial in the management of severe and, sometimes, intractable pain [147].

3.8.2. Histamine H4 Receptor (H4R)

H4R is expressed in a subset of small-to-medium-diameter sensory neurons and the spinal cord. Pharmacological blockade and the genetic deletion of H4R suggest that the activation of H4R triggers downstream itch signaling [148,149]. Evidence suggests that the analgesic effects of H4R manipulation are dependent on the administration routes. When various H4R antagonists were administered intrathecally in neuropathic pain models, pronociceptive effects were observed. These results are supported by the observation that the intrathecal administration of H4R agonists (ST-1006 and VUF8430) attenuates neuropathic pain [146]. By contrast, the intraperitoneal or subcutaneous administration of H4R antagonists, including TR-7 and JNJ7777120, alleviates pain in nerve injuries or inflammation in murine models [150,151]. Although JNJ7777120 displayed both analgesic and antipruritic effects in an experimental setting, its short half-life prevented its entry into clinical development. Nevertheless, this peripheral blockade approach for H4R seems promising for treating pain and itch at the same time.

3.9. Cannabinoid Receptors

Cannabinoid receptors belong to the G protein-coupled receptor family and are involved in multiple physiological processes, such as inflammation, cognition, itch, and pain. There are two identified receptors, cannabinoid receptor 1 (CB1) and cannabinoid receptor 2 (CB2), with the former mediating the modulatory effect in both the central and peripheral nervous system, and the latter only in the peripheral nervous system [152,153,154]. CB1 and CB2 share 44% of their sequence homology and are coupled with Gi/o downstream inhibitory signaling pathways [155].
CB1 and CB2 agonists may mediate antinociceptive and antipruritic effects, and a wide range of compounds have been tested for such effects [156]. The primary effects on pain and itch take place mainly by mediating the activities on the sensory neurons. These effects are partly produced by mediating the inflammatory process. Aside from the nervous system, CB1 and CB2 are also expressed in mast cells, macrophages, and the keratinocytes in the skin [157]. Therefore, some of the effects are mediated by the modification of immune cells and barriers in the skin.
The agonists of CB1 have been employed in the treatment of both pain and itch. However, due to disturbances in physiological functions caused by its activities on the central nervous system, CB1 agonists may bring about major side effects, such as hallucination and panic [158]. Given that CB2 is mainly distributed in the peripheral tissues, the administration of CB2-selective agonists may theoretically circumvent these side effects through CB1 activation. However, to date, most of the compounds with antinociceptive and antipruritic properties in humans have been found to be nonselective to CB2. One example is dronabinol (delta-9-tetrahydrocannabinol), a synthetic CB1 and CB2 agonist, which has been efficacious in treating intractable itch and chronic pain [159,160].
One practical strategy for avoiding the adverse effects of systemic cannabinoid agonists is via topical administration. However, although the topical administration of various cannabinoid agonists, including HU-210 and anandamide, has been reported for the treatment of itch [161,162,163], whether similar topical approaches exert antinociceptive effects in human remains unclear. Comprehensive investigations are warranted.

3.10. Oncostatin M (OSM)

Oncostatin M is a member of the IL-6 family and is a cytokine involved in organ development, inflammation, itch, and pain. Its receptor, oncostatin M receptor (OSMR), is expressed broadly in neural and non-neural cells. OSMR was initially implicated as a receptor for pain. In the sensory neuron, OSMR overlaps with TRPV1 and purinergic receptor P2X3 as a subset of nociceptive neurons [164]. OSMR-knockout mice displayed attenuated pain behaviors [165]. In the murine model, the administration of OSM enhanced hyperalgesia via prolonged ERK signaling and, eventually, led to the priming of chronic pain [166]. Recent investigations on OSMR reveal that it plays more important roles in itch sensitization. The itch intensity of chronic prurigo was found to be correlated to the level of OSM and OSMR-expressing cells in the skin [167]. A study by Tseng and Hoon showed that OSMR is responsible for itch sensitization, and OSM is upregulated in various inflammatory dermatoses and cutaneous T-cell lymphoma [168]. Nevertheless, currently, there is no commercially available agent targeting OSM-OSMR in the management of both pain and itch. In line with the theory that OSM-OSMR is involved in the pathogenesis of both pain and itch, the pharmacological blockade of this axis may lead to the treatment of both sensations in the future.

3.11. JAK-STAT (Janus Kinase/Signal Transducer and Activator of Transcription) Pathway

The JAK-STAT pathway was first identified in 1992. Over the past decade, substantial advances in the use of JAK-STAT inhibitors for treating various disorders have been made [169,170]. Currently, four members of the JAK family have been identified, namely JAK1, JAK2, JAK3, and TYK2, with seven members in the STAT family, including STAT1, STAT2, STAT3, STAT4, STAT5a, STAT5b, and STAT6 [171]. JAK-STAT signaling has been implicated widely in different physiological functions, including immunity, cell division, and cancer, as well as in pain and itch. Notably, there is shared JAK-STAT signaling in the neural and immune pathways, contributing to its roles in pain and itch. Key mediators in this process include IL-4, IL-13, Th1, Th2, and Th17. Moreover, both JAK1 and JAK2 have been found to be directly involved in the neural pathway for itch [172,173,174]. JAK-STAT inhibitors, including abrocitinib (JAK1 inhibitor), baricitinib (JAK1 and JAK2 inhibitors), ruxolitinib (JAK1 and JAK2 inhibitors), and upadacitinib (JAK1 inhibitor), can inhibit itch rather rapidly [175]. Moreover, mounting evidence from both basic and clinical studies indicates that JAK-STAT3 signaling plays a role in nociception. JAK-STAT3 signaling is crucial for microglia and astrocyte activation following nerve injuries, which subsequently contributes to pain [176,177]. Another proposed mechanism is that JAK-STAT inhibitors exert their antinociceptive functions by modulating pain-associated mediators such as IL-17, IL-6, and IL-23 [178]. Clinically, baricitinib and upadacitinib have been reported to alleviate pain and associated behaviors in rheumatoid arthritis patients [179].

3.12. Nerve Growth Factor (NGF)

NGF belongs to the neurotrophin family and is critical for the growth and maintenance of neurons. There are two receptors for NGF, including a high-affinity tyrosine kinase receptor, tropomyosin kinase receptor A (TrkA), and a nonselective and low-affinity p75 pan-neurotrophin receptor (p75NTR). The binding of NGF to TrkA leads to the activation of Ras/Raf/mitogen-activated protein kinase (MAPK), phosphatidylinositol 3-kinase (PI3K)/Akt, and phospholipase C-γ (PLC-γ) pathways, thus promoting the survival of cells. However, the absence or decreased expression of TrkA can induce the activation of p75NTR, which may trigger the JNK pathway and induce cell apoptosis [180,181]. Studies have indicated that the ratio of TrkA and p75NTR expressed on cell surfaces contributes crucially to the determination of NGF on cells [182,183]. Besides nervous systems, NGF has also been reported to be produced and utilized by other cell types, such as epithelial cells, endothelial cells, mast cells, and B and T cells. In sum, NGF plays a substantial role in the crosstalk among the nervous system, immune system, and other cell types [184,185].
Elevated NGF levels have been implicated in pain and itch for various disorders. In chronic prurigo, atopic dermatitis, and allergic contact dermatitis, NGF overexpression is accompanied by the hyperinnervation of skin nerves, which may cause itch or pain in the lesional skin [51,52,186]. These findings indicate the potential of NGF as a target to treat pain and itch in several diseases. Multiple anti-NGF monoclonal antibodies, including tanezumab and fasinumab, have been developed for the treatment of osteoarthritic pain, low back pain, and pain from other diseases. In theory, this approach may attenuate chronic pain and itch. However, severe adverse effects, such as osteonecrosis, have been reported, which suggests great risk in the development of anti-NGF antibodies, and, thus, many of these developments were halted [187,188].

3.13. Protease-Activated Receptor 2 (PAR2)

The protease-activated receptor (PAR) is a G protein-coupled receptor family that requires the proteolytic cleavage of the extracellular N-terminal domain to reveal a new cleaved N-terminus, which can function as a ligand to activate the PARs [189]. There are four members of the PAR family: PAR1, PAR2, PAR3, and PAR4. Among them, PAR2 is widely discussed for its role in both neural and immune systems, as well as its ability in triggering pain and itch [190,191]. PAR2 can be cleaved and activated by trypsin and mast cell tryptase. Research indicates that proteases from specific pathogens, such as house dust mites and some bacteria, can also induce the activation of PAR2 [192]. The expression of PAR2 is broadly seen on various cell types, including neurons, epithelial cells, keratinocytes, mast cells, and neutrophils. This suggests the integral role PAR2 plays in diseases related to these cells.
The role of PAR2 in pain is well established. Conversely, there have been mixed results regarding the role of PAR2 in itch, as current evidence indicates that PAR2 mediates itch via keratinocytes, not directly via sensory neurons [193,194]. Studies have revealed that the activation of PAR2 may contribute to several diseases, such as airway and lung inflammation, arthritis, skin diseases, neurological disorders, and chronic pain [195,196,197]. Thus, PAR2 antagonists were developed to treat these conditions, as well as to treat pain and itch. Animal experiments have supported the premise that PAR2 is a mediator of cancer pain, while the overexpression of PAR2 in atopic dermatitis mouse models induces a severe itch response [198,199,200]. Although various PAR2 antagonists, including peptides, pepducins, small molecules, and antibodies, have been introduced, only some of them have been tested for treating pain and itch. For example, FSLLRY-NH2 was proven to inhibit neuropathic pain in rats with spinal cord injury, and reduce dermatophyte-associated in atopic dermatitis mouse models; a PAR2 antibody, MEDI0618, is undergoing a Phase I clinical trial, with an aim for the treatment of chronic pain in the future [194,201,202,203].

3.14. Other Agents with Analgesic and Antipruritic Effects

Multiple agents with both analgesic and antipruritic effects are available now and are commonly prescribed clinically. Although the targets and mechanisms for many of them are not fully understood, they are discussed as follows because they are of undeniable scientific and clinical values.

3.14.1. Botulinum Toxin

Botulinum toxin is a neurotoxin produced by the bacterium Clostridium botulinum. There are seven types of botulinum toxins, with type A and type B toxins being utilized medically. In the synapses, these toxins disrupt the functions of the SNARE complex and prevent the release of neurotransmitters and, thus, interrupt pain and itch signaling. Botulinum toxin A has been employed in the treatment of various pain conditions, such as migraine, trigeminal neuralgia, postherpetic neuralgia, diabetic neuropathy, occipital neuralgia, complex regional pain syndrome, etc. [204,205]. Botulinum toxin A may also alleviate itch in various disorders, including notalgia paresthetica, keloid, multiple sclerosis, brachioradial pruritus, and Fox–Fordyce disease [206,207,208,209,210,211].

3.14.2. Local Anesthetics

In addition to their known antinociceptive effects, topical anesthetics are most widely used for the treatment of itch. The topical application of lidocaine has been reported to be effective in the treatment of different pruritic disorders, including liver disorders, burns, and other disorders when combined with ketamine and amitriptyline [212,213,214]. Lidocaine is also available as an over-the-counter agent for topical usage. Topical pramoxine has also been found to be effective for the treatment of uremic pruritus [215].

3.14.3. Gapapentin/Pregabalin

Gapapentin and its successor, pregabalin, have been approved for the treatment of painful conditions such as postherpetic neuralgia, spinal cord injury, diabetic neuropathy, and fibromyalgia. The exact mechanisms of gabapentin and pregabalin are still unclear. It is proposed that these two drugs attenuate itch and pain by inhibiting the α2δ unit in voltage-gated calcium channels on sensory neurons and the spinal cord [216]. In the meantime, gabapentin and pregabalin have also been reported to be effective in treating itch for various conditions. These disorders include uremic pruritus, brachioradial pruritus, chronic prurigo, spinal cord injury, and pruritus of unknown origin. [217].

4. Conclusions

Pain and itch are distinct sensations yet share some overlapping neural pathways and key molecules in their pathogenesis. Clinically chronic pain and chronic itch may be present at the same time in various disorders. Here we reviewed the common grounds of pain and itch and, at the same time, druggable targets for developing agents to treat pain and itch. Clearly, heterogeneity exists in pain and itch with respect to their molecular signatures, their pathways from the skin to the nervous system, and even in their manifestations for various disorders. The complex nature thus renders their pathogenesis difficult to understand and the creation of pharmacological agents unlikely to fully attenuate the symptoms based on the pathophysiology. With respect to this, it may be reasonable not to focus solely on the symptoms of pain and itch per se but on the symptoms in specific diseases with due regard to the heterogeneity of their pathophysiology. Further investigation is warranted to dissect the differences and similarities in pain and itch pathways, including skin biopsies, functional imaging, and psychophysics, along with techniques such as mouse genetics and in vivo imaging. This will unveil the potential of novel agents with both antinociceptive and antipruritic properties.

Author Contributions

Conceptualization, H.-J.W., Q.T.T.P. and T.-F.T.; writing—original draft preparation, H.-J.W., Q.T.T.P. and C.-W.C.; writing—review and editing, T.-F.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors acknowledge the academic and science graphic illustration service provided by the TMU Office of Research and Development. The authors also acknowledge the graphic illustration provided by Wikimedia Commons with the following links: https://commons.wikimedia.org/wiki/File:Mast_cell.svg (accessed on 22 May 2022); https://commons.wikimedia.org/wiki/File:Neutrophil.svg (accessed on 22 May 2022); https://commons.wikimedia.org/wiki/File:Dendritic_cell.svg (accessed on 22 May 2022).

Conflicts of Interest

Tsen-fang Tsai has conducted clinical trials or received honoraria for serving as a consultant for AbbVie, Boehringer Ingelheim, Bristol-Myers Squibb, Celgene, Eli Lilly, Galderma, GSK-Stiefel, Janssen-Cilag, Leo-Pharma, Merck, Novartis, Pfizer Inc., and UCB Pharma. Hao-Jui Weng has conducted clinical and preclinical trials for Microbio and Oneness Biotech.

References

  1. Honma, M.; Kanai, Y.; Murotani, K.; Nomura, T.; Ito, K.; Imafuku, S. Itching and skin pain in real-life patients with plaque psoriasis: Baseline analysis of the ProLOGUE study. J. Dermatol. Sci. 2022, 105, 189–191. [Google Scholar] [CrossRef] [PubMed]
  2. Newton, L.; DeLozier, A.M.; Griffiths, P.C.; Hill, J.N.; Hudgens, S.; Symonds, T.; Gable, J.C.; Paik, J.; Wyrwich, K.W.; Eichenfield, L.F. Exploring content and psychometric validity of newly developed assessment tools for itch and skin pain in atopic dermatitis. J. Patient-Rep. Outcomes 2019, 3, 42. [Google Scholar] [CrossRef] [PubMed]
  3. Oaklander, A.L.; Bowsher, D.; Galer, B.; Haanpää, M.; Jensen, M.P. Herpes zoster itch: Preliminary epidemiologic data. J. Pain 2003, 4, 338–343. [Google Scholar] [CrossRef]
  4. Maarouf, M.; Kromenacker, B.; Capozza, K.cL.; Kempton, D.; Hendricks, A.; Tran, K.; Shi, V.Y. Pain and itch are dual burdens in atopic dermatitis. Dermatitis 2018, 29, 278–281. [Google Scholar] [CrossRef] [PubMed]
  5. Dawn, A.; Papoiu, A.; Chan, Y.; Rapp, S.; Rassette, N.; Yosipovitch, G. Itch characteristics in atopic dermatitis: Results of a web-based questionnaire. Br. J. Dermatol. 2009, 160, 642–644. [Google Scholar] [CrossRef] [PubMed]
  6. Van Wijck, A.J.; Aerssens, Y.R. Pain, itch, quality of life, and costs after herpes zoster. Pain Pract. 2017, 17, 738–746. [Google Scholar] [CrossRef] [PubMed]
  7. Sauver, J.L.S.; Warner, D.O.; Yawn, B.P.; Jacobson, D.J.; McGree, M.E.; Pankratz, J.J.; Melton, L.J., III; Roger, V.L.; Ebbert, J.O.; Rocca, W.A. Why patients visit their doctors: Assessing the most prevalent conditions in a defined American population. Mayo Clin. Proc. 2013, 88, 56–67. [Google Scholar] [CrossRef] [Green Version]
  8. Vos, T.; Allen, C.; Arora, M.; Barber, R.M.; Bhutta, Z.A.; Brown, A.; Carter, A.; Casey, D.C.; Charlson, F.J.; Chen, A.Z. Global, regional, and national incidence, prevalence, and years lived with disability for 310 diseases and injuries, 1990–2015: A systematic analysis for the Global Burden of Disease Study 2015. Lancet 2016, 388, 1545–1602. [Google Scholar] [CrossRef] [Green Version]
  9. Dalgard, F.; Svensson, Å.; Holm, J.Ø.; Sundby, J. Self-reported skin morbidity among adults: Associations with quality of life and general health in a Norwegian survey. J. Investig. Dermatol. Symp. Proc. 2004, 9, 120–125. [Google Scholar] [CrossRef] [Green Version]
  10. Boston, A.; Sharpe, L. The role of threat-expectancy in acute pain: Effects on attentional bias, coping strategy effectiveness and response to pain. Pain 2005, 119, 168–175. [Google Scholar] [CrossRef]
  11. Roelofs, J.; Peters, M.L.; Vlaeyen, J.W. Selective attention for pain-related information in healthy individuals: The role of pain and fear. Eur. J. Pain 2002, 6, 331–339. [Google Scholar] [CrossRef]
  12. Fortune, D.G.; Richards, H.L.; Corrin, A.; Taylor, R.J.; Griffiths, C.E.; Main, C.J. Attentional bias for psoriasis-specific and psychosocial threat in patients with psoriasis. J. Behav. Med. 2003, 26, 211–224. [Google Scholar] [CrossRef] [PubMed]
  13. Van Laarhoven, A.I.; Kraaimaat, F.W.; Wilder-Smith, O.; Evers, A.W. Role of attentional focus on bodily sensations in sensitivity to itch and pain. Acta Derm. Venereol. 2010, 90, 46–51. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Seminowicz, D.A.; Davis, K.D. Pain enhances functional connectivity of a brain network evoked by performance of a cognitive task. J. Neurophysiol. 2007, 97, 3651–3659. [Google Scholar] [CrossRef] [PubMed]
  15. McGowan, N.; Sharpe, L.; Refshauge, K.; Nicholas, M. The effect of attentional re-training and threat expectancy in response to acute pain. Pain 2009, 142, 101–107. [Google Scholar] [CrossRef]
  16. Sikand, P.; Shimada, S.G.; Green, B.G.; LaMotte, R.H. Similar itch and nociceptive sensations evoked by punctate cutaneous application of capsaicin, histamine and cowhage. Pain 2009, 144, 66–75. [Google Scholar] [CrossRef] [Green Version]
  17. Whitty, C.W.M. Substances producing pain and itch. J. Neurol. Neurosurg. Psychiatry 1964, 27, 483. [Google Scholar] [CrossRef] [Green Version]
  18. Davidson, S.; Giesler, G.J. The multiple pathways for itch and their interactions with pain. Trends Neurosci. 2010, 33, 550–558. [Google Scholar] [CrossRef] [Green Version]
  19. Simpson, J. Pain and the Neurosurgeon a Forty-Year Experience. J. Neurol. Neurosurg. Psychiatry 1970, 33, 129. [Google Scholar] [CrossRef] [Green Version]
  20. Clark, K. Use of cordotomy in the relief of intractable pain. Arch. Surg. 1961, 82, 440–442. [Google Scholar] [CrossRef]
  21. Vedantam, A.; Bruera, E.; Hess, K.R.; Dougherty, P.M.; Viswanathan, A. Somatotopy and organization of spinothalamic tracts in the human cervical spinal cord. Neurosurgery 2019, 84, E311–E317. [Google Scholar] [CrossRef] [PubMed]
  22. Hyndman, O.R.; Wolkin, J. Anterior chordotomy: Further observations on physiologic results and optimum manner of performance. Arch. Neurol. Psychiatry 1943, 50, 129–148. [Google Scholar] [CrossRef]
  23. Han, L.; Ma, C.; Liu, Q.; Weng, H.-J.; Cui, Y.; Tang, Z.; Kim, Y.; Nie, H.; Qu, L.; Patel, K.N. A subpopulation of nociceptors specifically linked to itch. Nat. Neurosci. 2013, 16, 174–182. [Google Scholar] [CrossRef] [PubMed]
  24. Jinks, S.L.; Carstens, E. Responses of superficial dorsal horn neurons to intradermal serotonin and other irritants: Comparison with scratching behavior. J. Neurophysiol. 2002, 87, 1280–1289. [Google Scholar] [CrossRef] [Green Version]
  25. Johanek, L.M.; Meyer, R.A.; Friedman, R.M.; Greenquist, K.W.; Shim, B.; Borzan, J.; Hartke, T.; LaMotte, R.H.; Ringkamp, M. A role for polymodal C-fiber afferents in nonhistaminergic itch. J. Neurosci. 2008, 28, 7659–7669. [Google Scholar] [CrossRef] [Green Version]
  26. LaMotte, R.H.; Dong, X.; Ringkamp, M. Sensory neurons and circuits mediating itch. Nat. Rev. Neurosci. 2014, 15, 19–31. [Google Scholar] [CrossRef] [Green Version]
  27. Ma, Q. Labeled lines meet and talk: Population coding of somatic sensations. J. Clin. Investig. 2010, 120, 3773–3778. [Google Scholar] [CrossRef] [Green Version]
  28. Patel, K.N.; Dong, X. An itch to be scratched. Neuron 2010, 68, 334–339. [Google Scholar] [CrossRef] [Green Version]
  29. Liu, T.; Ji, R.-R. New insights into the mechanisms of itch: Are pain and itch controlled by distinct mechanisms? Pflügers Arch. Eur. J. Physiol. 2013, 465, 1671–1685. [Google Scholar] [CrossRef]
  30. Wang, H.; Papoiu, A.; Coghill, R.; Patel, T.; Wang, N.; Yosipovitch, G. Ethnic differences in pain, itch and thermal detection in response to topical capsaicin: African Americans display a notably limited hyperalgesia and neurogenic inflammation. Br. J. Dermatol. 2010, 162, 1023–1029. [Google Scholar] [CrossRef]
  31. Moser, H.R.; Giesler, G.J., Jr. Characterization of pruriceptive trigeminothalamic tract neurons in rats. J. Neurophysiol. 2014, 111, 1574–1589. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Klein, A.; Carstens, M.I.; Carstens, E. Facial injections of pruritogens or algogens elicit distinct behavior responses in rats and excite overlapping populations of primary sensory and trigeminal subnucleus caudalis neurons. J. Neurophysiol. 2011, 106, 1078–1088. [Google Scholar] [CrossRef] [PubMed]
  33. McNeil, B.; Dong, X. Peripheral mechanisms of itch. Neurosci. Bull. 2012, 28, 100–110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Yosipovitch, G.; Carstens, E.; McGlone, F. Chronic itch and chronic pain: Analogous mechanisms. Pain 2007, 131, 4–7. [Google Scholar] [CrossRef]
  35. Van Laarhoven, A.; Kraaimaat, F.; Wilder-Smith, O.; Van de Kerkhof, P.; Cats, H.; Van Riel, P.; Evers, A. Generalized and symptom-specific sensitization of chronic itch and pain. J. Eur. Acad. Dermatol. Venereol. 2007, 21, 1187–1192. [Google Scholar] [CrossRef]
  36. Davidson, R.J.; Jackson, D.C.; Kalin, N.H. Emotion, plasticity, context, and regulation: Perspectives from affective neuroscience. Psychol. Bull. 2000, 126, 890. [Google Scholar] [CrossRef]
  37. Brosschot, J.F. Cognitive-emotional sensitization and somatic health complaints. Scand. J. Psychol. 2002, 43, 113–121. [Google Scholar] [CrossRef]
  38. Houtveen, J.H.; Rietveld, S.; de Geus, E.J. Exaggerated perception of normal physiological responses to stress and hypercapnia in young women with numerous functional somatic symptoms. J. Psychosom. Res. 2003, 55, 481–490. [Google Scholar] [CrossRef] [Green Version]
  39. Rietvelt, S.; Houtveen, J.H. Acquired sensitivity to relevant physiological activity in patients with chronic health problems. Behav. Res. Ther. 2004, 42, 137–153. [Google Scholar] [CrossRef]
  40. Van den Bergh, O.; Winters, W.; Devriese, S.; Van Diest, I. Learning subjective health complaints. Scand. J. Psychol. 2002, 43, 147–152. [Google Scholar] [CrossRef]
  41. Pisoni, R.L.; Wikström, B.; Elder, S.J.; Akizawa, T.; Asano, Y.; Keen, M.L.; Saran, R.; Mendelssohn, D.C.; Young, E.W.; Port, F.K. Pruritus in haemodialysis patients: International results from the Dialysis Outcomes and Practice Patterns Study (DOPPS). Nephrol. Dial. Transplant. 2006, 21, 3495–3505. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Giesecke, T.; Williams, D.A.; Harris, R.E.; Cupps, T.R.; Tian, X.; Tian, T.X.; Gracely, R.H.; Clauw, D.J. Subgrouping of fibromyalgia patients on the basis of pressure-pain thresholds and psychological factors. Arthritis Rheum. 2003, 48, 2916–2922. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Giesecke, T.; Gracely, R.H.; Grant, M.A.; Nachemson, A.; Petzke, F.; Williams, D.A.; Clauw, D.J. Evidence of augmented central pain processing in idiopathic chronic low back pain. Arthritis Rheum. 2004, 50, 613–623. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Ikoma, A.; Fartasch, M.; Heyer, G.; Miyachi, Y.; Handwerker, H.; Schmelz, M. Painful stimuli evoke itch in patients with chronic pruritus: Central sensitization for itch. Neurology 2004, 62, 212–217. [Google Scholar] [CrossRef] [PubMed]
  45. Thieme, K.; Rose, U.; Pinkpank, T.; Spies, C.; Turk, D.C.; Flor, H. Psychophysiological responses in patients with fibromyalgia syndrome. J. Psychosom. Res. 2006, 61, 671–679. [Google Scholar] [CrossRef]
  46. Staud, R.; Rodriguez, M.E. Mechanisms of disease: Pain in fibromyalgia syndrome. Nat. Clin. Pract. Rheumatol. 2006, 2, 90–98. [Google Scholar] [CrossRef]
  47. Kidd, B.; Urban, L. Mechanisms of inflammatory pain. Br. J. Anaesth. 2001, 87, 3–11. [Google Scholar] [CrossRef] [Green Version]
  48. Schmelz, M.; Schmidt, R.; Weidner, C.; Hilliges, M.; Torebjork, H.; Handwerker, H.O. Chemical response pattern of different classes of C-nociceptors to pruritogens and algogens. J. Neurophysiol. 2003, 89, 2441–2448. [Google Scholar] [CrossRef] [Green Version]
  49. Baral, P.; Mills, K.; Pinho-Ribeiro, F.A.; Chiu, I.M. Pain and itch: Beneficial or harmful to antimicrobial defense? Cell Host Microbe 2016, 19, 755–759. [Google Scholar] [CrossRef] [Green Version]
  50. Toyoda, M.; Nakamura, M.; Makino, T.; Hino, T.; Kagoura, M.; Morohashi, M. Nerve growth factor and substance P are useful plasma markers of disease activity in atopic dermatitis. Br. J. Dermatol. 2002, 147, 71–79. [Google Scholar] [CrossRef]
  51. Kinkelin, I.; Mötzing, S.; Koltzenburg, M.; Bröcker, E.-B. Increase in NGF content and nerve fiber sprouting in human allergic contact eczema. Cell Tissue Res. 2000, 302, 31–37. [Google Scholar] [CrossRef] [PubMed]
  52. Johansson, O.; Liang, Y.; Emtestam, L. Increased nerve growth factor-and tyrosine kinase A-like immunoreactivities in prurigo nodularis skin–an exploration of the cause of neurohyperplasia. Arch. Dermatol. Res. 2002, 293, 614–619. [Google Scholar] [CrossRef] [PubMed]
  53. Choi, J.-C.; Yang, J.-H.; Chang, S.-E.; Choi, J.-H. Pruritus and nerve growth factor in psoriasis. Korean J. Dermatol. 2005, 43, 769–773. [Google Scholar]
  54. Bohm-Starke, N.; Hilliges, M.; Falconer, C.; Rylander, E. Increased intraepithelial innervation in women with vulvar vestibulitis syndrome. Gynecol. Obstet. Investig. 1998, 46, 256–260. [Google Scholar] [CrossRef] [PubMed]
  55. Halvorson, K.G.; Kubota, K.; Sevcik, M.A.; Lindsay, T.H.; Sotillo, J.E.; Ghilardi, J.R.; Rosol, T.J.; Boustany, L.; Shelton, D.L.; Mantyh, P.W. A blocking antibody to nerve growth factor attenuates skeletal pain induced by prostate tumor cells growing in bone. Cancer Res. 2005, 65, 9426–9435. [Google Scholar] [CrossRef] [Green Version]
  56. Schmelz, M.; Mantyh, P.; Malfait, A.-M.; Farrar, J.; Yaksh, T.; Tive, L.; Viktrup, L. Nerve growth factor antibody for the treatment of osteoarthritis pain and chronic low-back pain: Mechanism of action in the context of efficacy and safety. Pain 2019, 160, 2210–2220. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Atanassoff, P.G.; Brull, S.J.; Zhang, J.; Greenquist, K.; Silverman, D.G.; Lamotte, R.H. Enhancement of experimental pruritus and mechanically evoked dysesthesiae with local anesthesia. Somatosens. Mot. Res. 1999, 16, 291–298. [Google Scholar] [CrossRef]
  58. LaMotte, R.H. Subpopulations of “nocifensor neurons” contributing to pain and allodynia, itch and alloknesis. APS J. 1992, 1, 115–126. [Google Scholar] [CrossRef]
  59. Yosipovitch, G.; Rosen, J.D.; Hashimoto, T. Itch: From mechanism to (novel) therapeutic approaches. J. Allergy Clin. Immunol. 2018, 142, 1375–1390. [Google Scholar] [CrossRef] [Green Version]
  60. Simone, D.A.; Alreja, M.; Lamotte, R.H. Psychophysical studies of the itch sensation and itchy skin (“alloknesis”) produced by intracutaneous injection of histamine. Somatosens. Mot. Res. 1991, 8, 271–279. [Google Scholar] [CrossRef]
  61. Torebjörk, H.; Schmelz, M.; Handwerker, H. Functional Properties of Human Cutaneous Nociceptors and Their Role in Pain and Hyperalgesia. In Neurobiology of Nociceptors; Oxford University Press: Oxford, UK, 1996; pp. 349–369. [Google Scholar]
  62. LaMotte, R.H.; Shain, C.N.; Simone, D.A.; Tsai, E. Neurogenic hyperalgesia: Psychophysical studies of underlying mechanisms. J. Neurophysiol. 1991, 66, 190–211. [Google Scholar] [CrossRef] [PubMed]
  63. Koltzenburg, M. Neural mechanisms of cutaneous nociceptive pain. Clin. J. Pain 2000, 16, S131–S138. [Google Scholar] [CrossRef] [PubMed]
  64. Torebjörk, H.; Lundberg, L.; LaMotte, R. Central changes in processing of mechanoreceptive input in capsaicin-induced secondary hyperalgesia in humans. J. Physiol. 1992, 448, 765–780. [Google Scholar] [CrossRef] [PubMed]
  65. Magerl, W.; Fuchs, P.N.; Meyer, R.A.; Treede, R.-D. Roles of capsaicin-insensitive nociceptors in cutaneous pain and secondary hyperalgesia. Brain 2001, 124, 1754–1764. [Google Scholar] [CrossRef]
  66. Ziegler, E.; Magerl, W.; Meyer, R.; Treede, R.-D. Secondary hyperalgesia to punctate mechanical stimuli: Central sensitization to A-fibre nociceptor input. Brain 1999, 122, 2245–2257. [Google Scholar] [CrossRef] [Green Version]
  67. Fuchs, P.N.; Campbell, J.N.; Meyer, R.A. Secondary hyperalgesia persists in capsaicin desensitized skin. Pain 2000, 84, 141–149. [Google Scholar] [CrossRef]
  68. Baron, R.; Schwarz, K.; Kleinert, A.; Schattschneider, J.; Wasner, G. Histamine-induced itch converts into pain in neuropathic hyperalgesia. Neuroreport 2001, 12, 3475–3478. [Google Scholar] [CrossRef]
  69. Birklein, F.; Claus, D.; Riedl, B.; Neundörfer, B.; Handwerker, H.O. Effects of cutaneous histamine application in patients with sympathetic reflex dystrophy. Muscle Nerve 1997, 20, 1389–1395. [Google Scholar] [CrossRef]
  70. Kwatra, S.G.; Stander, S.; Bernhard, J.D.; Weisshaar, E.; Yosipovitch, G. Brachioradial pruritus: A trigger for generalization of itch. J. Am. Acad. Dermatol. 2013, 68, 870–873. [Google Scholar] [CrossRef]
  71. Hosogi, M.; Schmelz, M.; Miyachi, Y.; Ikoma, A. Bradykinin is a potent pruritogen in atopic dermatitis: A switch from pain to itch. Pain 2006, 126, 16–23. [Google Scholar] [CrossRef] [Green Version]
  72. Ishiuji, Y.; Coghill, R.; Patel, T.; Dawn, A.; Fountain, J.; Oshiro, Y.; Yosipovitch, G. Repetitive scratching and noxious heat do not inhibit histamine-induced itch in atopic dermatitis. Br. J. Dermatol. 2008, 158, 78–83. [Google Scholar] [CrossRef] [PubMed]
  73. Nilsson, H.-J.; Schouenborg, J. Differential inhibitory effect on human nociceptive skin senses induced by local stimulation of thin cutaneous fibers. Pain 1999, 80, 103–112. [Google Scholar] [CrossRef]
  74. Oaklander, A.L.; Cohen, S.P.; Raju, S.V. Intractable postherpetic itch and cutaneous deafferentation after facial shingles. Pain 2002, 96, 9–12. [Google Scholar] [CrossRef]
  75. Ishikawa, R.; Iseki, M.; Koga, R.; Yamaguchi, K.; Inada, E. Investigation of neuropathic pain component by the stage of a disease of the herpes zoster associated pain patients received pain clinic treatments. Pain Res. 2016, 31, 156–165. [Google Scholar] [CrossRef] [Green Version]
  76. Bueller, H.A. Gabapentin treatment for brachioradial pruritus. J. Eur. Acad. Derm. Venereol. 1999, 13, 227–230. [Google Scholar] [CrossRef]
  77. Caterina, M.J.; Schumacher, M.A.; Tominaga, M.; Rosen, T.A.; Levine, J.D.; Julius, D. The capsaicin receptor: A heat-activated ion channel in the pain pathway. Nature 1997, 389, 816–824. [Google Scholar] [CrossRef]
  78. Tominaga, M.; Caterina, M.J.; Malmberg, A.B.; Rosen, T.A.; Gilbert, H.; Skinner, K.; Raumann, B.E.; Basbaum, A.I.; Julius, D. The cloned capsaicin receptor integrates multiple pain-producing stimuli. Neuron 1998, 21, 531–543. [Google Scholar] [CrossRef] [Green Version]
  79. Caterina, M.J.; Leffler, A.; Malmberg, A.B.; Martin, W.; Trafton, J.; Petersen-Zeitz, K.; Koltzenburg, M.; Basbaum, A.; Julius, D. Impaired nociception and pain sensation in mice lacking the capsaicin receptor. Science 2000, 288, 306–313. [Google Scholar] [CrossRef]
  80. Wooten, M.; Weng, H.-J.; Hartke, T.V.; Borzan, J.; Klein, A.H.; Turnquist, B.; Dong, X.; Meyer, R.A.; Ringkamp, M. Three functionally distinct classes of C-fibre nociceptors in primates. Nat. Commun. 2014, 5, 4122. [Google Scholar] [CrossRef] [Green Version]
  81. Imamachi, N.; Park, G.H.; Lee, H.; Anderson, D.J.; Simon, M.I.; Basbaum, A.I.; Han, S.-K. TRPV1-expressing primary afferents generate behavioral responses to pruritogens via multiple mechanisms. Proc. Natl. Acad. Sci. USA 2009, 106, 11330–11335. [Google Scholar] [CrossRef] [Green Version]
  82. Lim, K.-M.; Park, Y.-H. Development of PAC-14028, a novel transient receptor potential vanilloid type 1 (TRPV1) channel antagonist as a new drug for refractory skin diseases. Arch. Pharmacal Res. 2012, 35, 393–396. [Google Scholar] [CrossRef] [PubMed]
  83. Yun, J.-W.; Seo, J.A.; Jang, W.-H.; Koh, H.J.; Bae, I.-H.; Park, Y.-H.; Lim, K.-M. Antipruritic effects of TRPV1 antagonist in murine atopic dermatitis and itching models. J. Investig. Dermatol. 2011, 131, 1576–1579. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Park, C.W.; Kim, B.J.; Lee, Y.W.; Won, C.; Park, C.O.; Chung, B.Y.; Lee, D.H.; Jung, K.; Nam, H.-J.; Choi, G. Asivatrep, a TRPV1 antagonist, for the topical treatment of atopic dermatitis: Phase 3, randomized, vehicle-controlled study (CAPTAIN-AD). J. Allergy Clin. Immunol. 2022, 149, 1340–1347.e1344. [Google Scholar] [CrossRef] [PubMed]
  85. Nikolaeva-Koleva, M.; Butron, L.; González-Rodríguez, S.; Devesa, I.; Valente, P.; Serafini, M.; Genazzani, A.A.; Pirali, T.; Ballester, G.F.; Fernández-Carvajal, A. A capsaicinoid-based soft drug, AG1529, for attenuating TRPV1-mediated histaminergic and inflammatory sensory neuron excitability. Sci. Rep. 2021, 11, 246. [Google Scholar] [CrossRef]
  86. Backonja, M.M.; Malan, T.P.; Vanhove, G.F.; Tobias, J.K. NGX-4010, a high-concentration capsaicin patch, for the treatment of postherpetic neuralgia: A randomized, double-blind, controlled study with an open-label extension. Pain Med. 2010, 11, 600–608. [Google Scholar] [CrossRef] [Green Version]
  87. Papoiu, A.D.; Yosipovitch, G. Topical capsaicin. The fire of a ‘hot’medicine is reignited. Expert Opin. Pharmacother. 2010, 11, 1359–1371. [Google Scholar] [CrossRef]
  88. Andersen, H.H.; Arendt-Nielsen, L.; Elberling, J. Topical capsaicin 8% for the treatment of neuropathic itch conditions. Clin. Exp. Dermatol. 2017, 42, 596–598. [Google Scholar] [CrossRef]
  89. Fazio, S.B.; Yosipovitch, G. Pruritus: Therapies for Localized Pruritus; Dellavalle, R.P., Callen, J., Eds.; Uptodate: Waltham, MA, USA, 2022. [Google Scholar]
  90. Smith, G.; Gunthorpe, M.; Kelsell, R.; Hayes, P.; Reilly, P.; Facer, P.; Wright, J.; Jerman, J.; Walhin, J.-P.; Ooi, L. TRPV3 is a temperature-sensitive vanilloid receptor-like protein. Nature 2002, 418, 186–190. [Google Scholar] [CrossRef]
  91. Moqrich, A.; Hwang, S.W.; Earley, T.J.; Petrus, M.J.; Murray, A.N.; Spencer, K.S.; Andahazy, M.; Story, G.M.; Patapoutian, A. Impaired thermosensation in mice lacking TRPV3, a heat and camphor sensor in the skin. Science 2005, 307, 1468–1472. [Google Scholar] [CrossRef]
  92. Yoshioka, T.; Imura, K.; Asakawa, M.; Suzuki, M.; Oshima, I.; Hirasawa, T.; Sakata, T.; Horikawa, T.; Arimura, A. Impact of the Gly573Ser substitution in TRPV3 on the development of allergic and pruritic dermatitis in mice. J. Investig. Dermatol. 2009, 129, 714–722. [Google Scholar] [CrossRef] [Green Version]
  93. Han, Y.; Luo, A.; Kamau, P.M.; Takomthong, P.; Hu, J.; Boonyarat, C.; Luo, L.; Lai, R. A plant-derived TRPV3 inhibitor suppresses pain and itch. Br. J. Pharmacol. 2021, 178, 1669–1683. [Google Scholar] [CrossRef] [PubMed]
  94. Liu, Q.; Wang, J.; Wei, X.; Hu, J.; Ping, C.; Gao, Y.; Xie, C.; Wang, P.; Cao, P.; Cao, Z. Therapeutic inhibition of keratinocyte TRPV3 sensory channel by local anesthetic dyclonine. eLife 2021, 10, e68128. [Google Scholar] [CrossRef] [PubMed]
  95. Shelmire, B.; Gastineau, F.; Shields, T.L. Evaluation of a new topical anesthetic, dyclonine hydrochloride. AMA Arch. Dermatol. 1955, 71, 728–730. [Google Scholar] [CrossRef] [PubMed]
  96. Broad, L.M.; Mogg, A.J.; Eberle, E.; Tolley, M.; Li, D.L.; Knopp, K.L. TRPV3 in drug development. Pharmaceuticals 2016, 9, 55. [Google Scholar] [CrossRef] [Green Version]
  97. Grubisha, O.; Mogg, A.J.; Sorge, J.L.; Ball, L.J.; Sanger, H.; Ruble, C.L.; Folly, E.A.; Ursu, D.; Broad, L.M. Pharmacological profiling of the TRPV3 channel in recombinant and native assays. Br. J. Pharmacol. 2014, 171, 2631–2644. [Google Scholar] [CrossRef] [Green Version]
  98. Liedtke, W.; Choe, Y.; Martí-Renom, M.A.; Bell, A.M.; Denis, C.S.; Hudspeth, A.; Friedman, J.M.; Heller, S. Vanilloid receptor–related osmotically activated channel (VR-OAC), a candidate vertebrate osmoreceptor. Cell 2000, 103, 525–535. [Google Scholar] [CrossRef] [Green Version]
  99. Alessandri-Haber, N.; Joseph, E.; Dina, O.A.; Liedtke, W.; Levine, J.D. TRPV4 mediates pain-related behavior induced by mild hypertonic stimuli in the presence of inflammatory mediator. Pain 2005, 118, 70–79. [Google Scholar] [CrossRef]
  100. Chen, Y.; Kanju, P.; Fang, Q.; Lee, S.H.; Parekh, P.K.; Lee, W.; Moore, C.; Brenner, D.; Gereau, R.W., IV; Wang, F. TRPV4 is necessary for trigeminal irritant pain and functions as a cellular formalin receptor. Pain 2014, 155, 2662–2672. [Google Scholar] [CrossRef] [Green Version]
  101. Akiyama, T.; Ivanov, M.; Nagamine, M.; Davoodi, A.; Carstens, M.I.; Ikoma, A.; Cevikbas, F.; Kempkes, C.; Buddenkotte, J.; Steinhoff, M. Involvement of TRPV4 in serotonin-evoked scratching. J. Investig. Dermatol. 2016, 136, 154–160. [Google Scholar] [CrossRef] [Green Version]
  102. Kim, S.; Barry, D.M.; Liu, X.-Y.; Yin, S.; Munanairi, A.; Meng, Q.-T.; Cheng, W.; Mo, P.; Wan, L.; Liu, S.-B. Facilitation of TRPV4 by TRPV1 is required for itch transmission in some sensory neuron populations. Sci. Signal. 2016, 9, ra71. [Google Scholar] [CrossRef] [Green Version]
  103. Lawhorn, B.G.; Brnardic, E.J.; Behm, D.J. TRPV4 antagonists: A patent review (2015–2020). Expert Opin. Ther. Pat. 2021, 31, 773–784. [Google Scholar] [CrossRef] [PubMed]
  104. Story, G.M.; Peier, A.M.; Reeve, A.J.; Eid, S.R.; Mosbacher, J.; Hricik, T.R.; Earley, T.J.; Hergarden, A.C.; Andersson, D.A.; Hwang, S.W. ANKTM1, a TRP-like channel expressed in nociceptive neurons, is activated by cold temperatures. Cell 2003, 112, 819–829. [Google Scholar] [CrossRef] [Green Version]
  105. Liu, Q.; Feng, L.; Han, X.; Zhang, W.; Zhang, H.; Xu, L. The TRPA1 Channel Mediates Mechanical Allodynia and Thermal Hyperalgesia in a Rat Bone Cancer Pain Model. Front. Pain Res. 2021, 2, 7. [Google Scholar] [CrossRef]
  106. Bandell, M.; Story, G.M.; Hwang, S.W.; Viswanath, V.; Eid, S.R.; Petrus, M.J.; Earley, T.J.; Patapoutian, A. Noxious cold ion channel TRPA1 is activated by pungent compounds and bradykinin. Neuron 2004, 41, 849–857. [Google Scholar] [CrossRef] [Green Version]
  107. Kwan, K.Y.; Glazer, J.M.; Corey, D.P.; Rice, F.L.; Stucky, C.L. TRPA1 modulates mechanotransduction in cutaneous sensory neurons. J. Neurosci. 2009, 29, 4808–4819. [Google Scholar] [CrossRef] [Green Version]
  108. Wilson, S.R.; Gerhold, K.A.; Bifolck-Fisher, A.; Liu, Q.; Patel, K.N.; Dong, X.; Bautista, D.M. TRPA1 is required for histamine-independent, Mas-related G protein–coupled receptor–mediated itch. Nat. Neurosci. 2011, 14, 595–602. [Google Scholar] [CrossRef] [Green Version]
  109. Kremeyer, B.; Lopera, F.; Cox, J.J.; Momin, A.; Rugiero, F.; Marsh, S.; Woods, C.G.; Jones, N.G.; Paterson, K.J.; Fricker, F.R. A gain-of-function mutation in TRPA1 causes familial episodic pain syndrome. Neuron 2010, 66, 671–680. [Google Scholar] [CrossRef] [Green Version]
  110. Rosengaard, C.; Andersen, H.H.; Arendt-Nielsen, L.; Gazerani, P. A human surrogate model of itch utilizing the TRPA1 agonist trans-cinnamaldehyde. Acta Derm. 2015, 95, 798–803. [Google Scholar]
  111. Chen, H.; Terrett, J.A. Transient receptor potential ankyrin 1 (TRPA1) antagonists: A patent review (2015–2019). Expert Opin. Ther. Pat. 2020, 30, 643–657. [Google Scholar] [CrossRef]
  112. Weng, H.-J.; Patel, K.N.; Jeske, N.A.; Bierbower, S.M.; Zou, W.; Tiwari, V.; Zheng, Q.; Tang, Z.; Mo, G.C.; Wang, Y. Tmem100 is a regulator of TRPA1-TRPV1 complex and contributes to persistent pain. Neuron 2015, 85, 833–846. [Google Scholar] [CrossRef] [Green Version]
  113. Dong, X.; Weng, H.-J. Tmem100 Peptides and Variants Thereof and Their Use in Treating or Preventing Diseases or Conditions. US11066455B2, 20 July 2021. [Google Scholar]
  114. Dhaka, A.; Murray, A.N.; Mathur, J.; Earley, T.J.; Petrus, M.J.; Patapoutian, A. TRPM8 is required for cold sensation in mice. Neuron 2007, 54, 371–378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. McKemy, D.D.; Neuhausser, W.M.; Julius, D. Identification of a cold receptor reveals a general role for TRP channels in thermosensation. Nature 2002, 416, 52–58. [Google Scholar] [CrossRef] [PubMed]
  116. González-Muñiz, R.; Bonache, M.A.; Martín-Escura, C.; Gómez-Monterrey, I. Recent progress in TRPM8 modulation: An update. Int. J. Mol. Sci. 2019, 20, 2618. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Davies, S.J.; Harding, L.M.; Baranowski, A.P. A novel treatment of postherpetic neuralgia using peppermint oil. Clin. J. Pain 2002, 18, 200–202. [Google Scholar] [CrossRef]
  118. Pergolizzi, J., Jr.; Taylor, R., Jr.; LeQuang, J.A.; Raffa, R.; Group, N.R. The role and mechanism of action of menthol in topical analgesic products. J. Clin. Pharm. Ther. 2018, 43, 313–319. [Google Scholar] [CrossRef] [Green Version]
  119. Caceres, A.I.; Liu, B.; Jabba, S.V.; Achanta, S.; Morris, J.B.; Jordt, S.E. Transient receptor potential cation channel subfamily M member 8 channels mediate the anti-inflammatory effects of eucalyptol. Br. J. Pharmacol. 2017, 174, 867–879. [Google Scholar] [CrossRef]
  120. Knowlton, W.M.; Palkar, R.; Lippoldt, E.K.; McCoy, D.D.; Baluch, F.; Chen, J.; McKemy, D.D. A sensory-labeled line for cold: TRPM8-expressing sensory neurons define the cellular basis for cold, cold pain, and cooling-mediated analgesia. J. Neurosci. 2013, 33, 2837–2848. [Google Scholar] [CrossRef] [Green Version]
  121. Andersen, H.H.; Gazerani, P.; Arendt-Nielsen, L. High-concentration L-menthol exhibits counter-irritancy to neurogenic inflammation, thermal and mechanical hyperalgesia caused by trans-cinnamaldehyde. J. Pain 2016, 17, 919–929. [Google Scholar] [CrossRef] [Green Version]
  122. Colvin, L.A.; Johnson, P.R.; Mitchell, R.; Fleetwood-Walker, S.M.; Fallon, M. From bench to bedside: A case of rapid reversal of bortezomib-induced neuropathic pain by the TRPM8 activator, menthol. J. Clin. Oncol. 2008, 26, 4519–4520. [Google Scholar] [CrossRef]
  123. Palkar, R.; Ongun, S.; Catich, E.; Li, N.; Borad, N.; Sarkisian, A.; McKemy, D.D. Cooling relief of acute and chronic itch requires TRPM8 channels and neurons. J. Investig. Dermatol. 2018, 138, 1391–1399. [Google Scholar] [CrossRef] [Green Version]
  124. Wei, E.T. Di-Isopropyl-Phosphinoyl-Alkanes (Dapa) Compounds as Topical Agents for the Treatment of Sensory Discomfort. US20210401857A1, 30 December 2021. [Google Scholar]
  125. Liu, B.; Fan, L.; Balakrishna, S.; Sui, A.; Morris, J.B.; Jordt, S.-E. TRPM8 is the principal mediator of menthol-induced analgesia of acute and inflammatory pain. Pain 2013, 154, 2169–2177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Liu, Y.; Liu, Y.; Narang, C.; Limjunyawong, N.; Jamaldeen, H.; Yu, S.; Patiram, S.; Nie, H.; Caterina, M.J.; Dong, X. Sensory Neuron Expressed TRPC3 Mediates Acute and Chronic Itch. 2021. Available online: https://assets.researchsquare.com/files/rs-1021582/v1/e34a7342-53a2-4e30-9468-4fa4759b6152.pdf?c=1648558243 (accessed on 4 June 2022).
  127. Ko, M.-C. Neuraxial Opioid-Induced Itch and Its Pharmacological Antagonism. In Pharmacology of Itch; Springer: Berlin/Heidelberg, Germany, 2015; pp. 315–335. [Google Scholar]
  128. Ko, M.H.; Song, M.; Edwards, T.; Lee, H.; Naughton, N. The role of central μ opioid receptors in opioid-induced itch in primates. J. Pharmacol. Exp. Ther. 2004, 310, 169–176. [Google Scholar] [CrossRef] [PubMed]
  129. Bigliardi, P.L.; Stammer, H.; Jost, G.; Rufli, T.; Büchner, S.; Bigliardi-Qi, M. Treatment of pruritus with topically applied opiate receptor antagonist. J. Am. Acad. Dermatol. 2007, 56, 979–988. [Google Scholar] [CrossRef]
  130. Fishbane, S.; Jamal, A.; Munera, C.; Wen, W.; Menzaghi, F. A phase 3 trial of difelikefalin in hemodialysis patients with pruritus. N. Engl. J. Med. 2020, 382, 222–232. [Google Scholar] [CrossRef]
  131. Menzaghi, F.; Spencer, R.; Abrouk, N.; Lewis, M.; Chalmers, D. (422) CR845, a peripheral kappa opioid, provides better pain relief with less nausea and vomiting than placebo in patients after bunionectomy. J. Pain 2015, 16, S81. [Google Scholar] [CrossRef]
  132. Wang, X.; Gou, X.; Yu, X.; Bai, D.; Tan, B.; Cao, P.; Qian, M.; Zheng, X.; Wang, H.; Tang, P. Antinociceptive and antipruritic effects of HSK21542, a peripherally-restricted kappa opioid receptor agonist, in animal models of pain and itch. Front. Pharmacol. 2021, 12, 773204. [Google Scholar] [CrossRef] [PubMed]
  133. Brust, T.F.; Morgenweck, J.; Kim, S.A.; Rose, J.H.; Locke, J.L.; Schmid, C.L.; Zhou, L.; Stahl, E.L.; Cameron, M.D.; Scarry, S.M. Biased agonists of the kappa opioid receptor suppress pain and itch without causing sedation or dysphoria. Sci. Signal. 2016, 9, ra117. [Google Scholar] [CrossRef] [Green Version]
  134. Lee, B.; Elston, D.M. The uses of naltrexone in dermatologic conditions. J. Am. Acad. Dermatol. 2019, 80, 1746–1752. [Google Scholar] [CrossRef]
  135. Lee, J.; Shin, J.U.; Noh, S.; Park, C.O.; Lee, K.H. Clinical efficacy and safety of naltrexone combination therapy in older patients with severe pruritus. Ann. Dermatol. 2016, 28, 159–163. [Google Scholar] [CrossRef] [Green Version]
  136. Levine, J.; Gordon, N.; Jones, R.; Fields, H. The narcotic antagonist naloxone enhances clinical pain. Nature 1978, 272, 826–827. [Google Scholar] [CrossRef]
  137. Chen, K.Y.; Chen, L.; Mao, J. Buprenorphine–naloxone therapy in pain management. Anesthesiology 2014, 120, 1262–1274. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Dawn, A.G.; Yosipovitch, G. Butorphanol for treatment of intractable pruritus. J. Am. Acad. Dermatol. 2006, 54, 527–531. [Google Scholar] [CrossRef] [PubMed]
  139. Golpanian, R.S.; Yosipovitch, G.; Levy, C. Use of butorphanol as treatment for cholestatic itch. Dig. Dis. Sci. 2021, 66, 1693–1699. [Google Scholar] [CrossRef] [PubMed]
  140. Mathur, V.S.; Kumar, J.; Crawford, P.W.; Hait, H.; Sciascia, T.; Investigators, T.S. A multicenter, randomized, double-blind, placebo-controlled trial of nalbuphine ER tablets for uremic pruritus. Am. J. Nephrol. 2017, 46, 450–458. [Google Scholar] [CrossRef]
  141. Jannuzzi, R.G. Nalbuphine for treatment of opioid-induced pruritus. Clin. J. Pain 2016, 32, 87–93. [Google Scholar] [CrossRef]
  142. Pereira, M.P.; Ständer, S. Novel drugs for the treatment of chronic pruritus. Expert Opin. Investig. Drugs 2018, 27, 981–988. [Google Scholar] [CrossRef]
  143. Thangam, E.B.; Jemima, E.A.; Singh, H.; Baig, M.S.; Khan, M.; Mathias, C.B.; Church, M.K.; Saluja, R. The role of histamine and histamine receptors in mast cell-mediated allergy and inflammation: The hunt for new therapeutic targets. Front. Immunol. 2018, 9, 1873. [Google Scholar] [CrossRef] [Green Version]
  144. Lewis, T. The blood vessels of the human skin. Br. Med. J. 1926, 2, 61. [Google Scholar] [CrossRef] [Green Version]
  145. Simone, D.A.; Ngeow, J.Y.; Whitehouse, J.; Becerra-Cabal, L.; Putterman, G.J.; Lamotte, R.H. The magnitude and duration of itch produced by intracutaneous injections of histamine. Somatosens. Res. 1987, 5, 81–92. [Google Scholar] [CrossRef]
  146. Obara, I.; Telezhkin, V.; Alrashdi, I.; Chazot, P.L. Histamine, histamine receptors, and neuropathic pain relief. Br. J. Pharmacol. 2020, 177, 580–599. [Google Scholar] [CrossRef]
  147. Santiago-Palma, J.; Fischberg, D.; Kornick, C.; Khjainova, N.; Gonzales, G. Diphenhydramine as an analgesic adjuvant in refractory cancer pain. J. Pain Symptom Manag. 2001, 22, 699–703. [Google Scholar] [CrossRef]
  148. Dunford, P.J.; Williams, K.N.; Desai, P.J.; Karlsson, L.; McQueen, D.; Thurmond, R.L. Histamine H4 receptor antagonists are superior to traditional antihistamines in the attenuation of experimental pruritus. J. Allergy Clin. Immunol. 2007, 119, 176–183. [Google Scholar] [CrossRef] [PubMed]
  149. Roßbach, K.; Wendorff, S.; Sander, K.; Stark, H.; Gutzmer, R.; Werfel, T.; Kietzmann, M.; Bäumer, W. Histamine H4 receptor antagonism reduces hapten-induced scratching behaviour but not inflammation. Exp. Dermatol. 2009, 18, 57–63. [Google Scholar] [CrossRef]
  150. Coruzzi, G.; Adami, M.; Guaita, E.; de Esch, I.J.; Leurs, R. Antiinflammatory and antinociceptive effects of the selective histamine H4-receptor antagonists JNJ7777120 and VUF6002 in a rat model of carrageenan-induced acute inflammation. Eur. J. Pharmacol. 2007, 563, 240–244. [Google Scholar] [CrossRef] [PubMed]
  151. Popiolek-Barczyk, K.; Łażewska, D.; Latacz, G.; Olejarz, A.; Makuch, W.; Stark, H.; Kieć-Kononowicz, K.; Mika, J. Antinociceptive effects of novel histamine H3 and H4 receptor antagonists and their influence on morphine analgesia of neuropathic pain in the mouse. Br. J. Pharmacol. 2018, 175, 2897–2910. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Schlosburg, J.E.; O’Neal, S.T.; Conrad, D.H.; Lichtman, A.H. CB1 receptors mediate rimonabant-induced pruritic responses in mice: Investigation of locus of action. Psychopharmacology 2011, 216, 323–331. [Google Scholar] [CrossRef] [Green Version]
  153. Clayton, N.; Marshall, F.; Bountra, C.; O’shaughnessy, C. CB1 and CB2 cannabinoid receptors are implicated in inflammatory pain. Pain 2002, 96, 253–260. [Google Scholar] [CrossRef]
  154. Bilir, K.; Anli, G.; Ozkan, E.; Gunduz, O.; Ulugol, A. Involvement of spinal cannabinoid receptors in the antipruritic effects of WIN 55,212-2, a cannabinoid receptor agonist. Clin. Exp. Dermatol. 2018, 43, 553–558. [Google Scholar] [CrossRef]
  155. Latek, D.; Kolinski, M.; Ghoshdastider, U.; Debinski, A.; Bombolewski, R.; Plazinska, A.; Jozwiak, K.; Filipek, S. Modeling of ligand binding to G protein coupled receptors: Cannabinoid CB1, CB2 and adrenergic β2AR. J. Mol. Model. 2011, 17, 2353–2366. [Google Scholar] [CrossRef]
  156. Avila, C.; Massick, S.; Kaffenberger, B.H.; Kwatra, S.G.; Bechtel, M. Cannabinoids for the treatment of chronic pruritus: A review. J. Am. Acad. Dermatol. 2020, 82, 1205–1212. [Google Scholar] [CrossRef]
  157. Ständer, S.; Schmelz, M.; Metze, D.; Luger, T.; Rukwied, R. Distribution of cannabinoid receptor 1 (CB1) and 2 (CB2) on sensory nerve fibers and adnexal structures in human skin. J. Dermatol. Sci. 2005, 38, 177–188. [Google Scholar] [CrossRef]
  158. Moreira, F.A.; Grieb, M.; Lutz, B. Central side-effects of therapies based on CB1 cannabinoid receptor agonists and antagonists: Focus on anxiety and depression. Best Pract. Res. Clin. Endocrinol. Metab. 2009, 23, 133–144. [Google Scholar] [CrossRef]
  159. Neff, G.W.; O’Brien, C.B.; Reddy, K.R.; Bergasa, N.V.; Regev, A.; Molina, E.; Amaro, R.; Rodriguez, M.J.; Chase, V.; Jeffers, L. Preliminary observation with dronabinol in patients with intractable pruritus secondary to cholestatic liver disease. Am. J. Gastroenterol. 2002, 97, 2117–2119. [Google Scholar] [CrossRef]
  160. Narang, S.; Gibson, D.; Wasan, A.D.; Ross, E.L.; Michna, E.; Nedeljkovic, S.S.; Jamison, R.N. Efficacy of dronabinol as an adjuvant treatment for chronic pain patients on opioid therapy. J. Pain 2008, 9, 254–264. [Google Scholar] [CrossRef]
  161. Dvorak, M.; Watkinson, A.; McGlone, F.; Rukwied, R. Histamine induced responses are attenuated by a cannabinoid receptor agonist in human skin. Inflamm. Res. 2003, 52, 238–245. [Google Scholar] [CrossRef]
  162. Yuan, C.; Wang, X.-M.; Guichard, A.; Tan, Y.-M.; Qian, C.-Y.; Yang, L.-J.; Humbert, P. N-palmitoylethanolamine and N-acetylethanolamine are effective in asteatotic eczema: Results of a randomized, double-blind, controlled study in 60 patients. Clin. Interv. Aging 2014, 9, 1163. [Google Scholar] [CrossRef] [Green Version]
  163. Szepietowski, J.C.; Szepietowski, T.; Reich, A. Efficacy and tolerance of the cream containing structured physiological lipids with endocannabinoids in the treatment of uremic pruritus: A preliminary study. Acta Dermatovenerol. Croat. 2005, 13, 97–103. [Google Scholar]
  164. Tamura, S.; Morikawa, Y.; Miyajima, A.; Senba, E. Expression of oncostatin M receptor β in a specific subset of nociceptive sensory neurons. Eur. J. Neurosci. 2003, 17, 2287–2298. [Google Scholar] [CrossRef]
  165. Morikawa, Y.; Tamura, S.; Minehata, K.-I.; Donovan, P.J.; Miyajima, A.; Senba, E. Essential function of oncostatin m in nociceptive neurons of dorsal root ganglia. J. Neurosci. 2004, 24, 1941–1947. [Google Scholar] [CrossRef] [Green Version]
  166. Garza Carbajal, A.; Ebersberger, A.; Thiel, A.; Ferrari, L.; Acuna, J.; Brosig, S.; Isensee, J.; Moeller, K.; Siobal, M.; Rose-John, S. Oncostatin M induces hyperalgesic priming and amplifies signaling of cAMP to ERK by RapGEF2 and PKA. J. Neurochem. 2021, 157, 1821–1837. [Google Scholar] [CrossRef]
  167. Hashimoto, T.; Nattkemper, L.A.; Kim, H.S.; Kursewicz, C.D.; Fowler, E.; Shah, S.M.; Nanda, S.; Fayne, R.A.; Paolini, J.F.; Romanelli, P. Itch intensity in prurigo nodularis is closely related to dermal interleukin-31, oncostatin M, IL-31 receptor alpha and oncostatin M receptor beta. Exp. Dermatol. 2021, 30, 804–810. [Google Scholar] [CrossRef]
  168. Tseng, P.-Y.; Hoon, M.A. Oncostatin M can sensitize sensory neurons in inflammatory pruritus. Sci. Transl. Med. 2021, 13, eabe3037. [Google Scholar] [CrossRef]
  169. Fu, X.-Y.; Kessler, D.S.; Veals, S.A.; Levy, D.E.; Darnell, J. ISGF3, the transcriptional activator induced by interferon alpha, consists of multiple interacting polypeptide chains. Proc. Natl. Acad. Sci. USA 1990, 87, 8555–8559. [Google Scholar] [CrossRef] [Green Version]
  170. Wilks, A.F. Two putative protein-tyrosine kinases identified by application of the polymerase chain reaction. Proc. Natl. Acad. Sci. USA 1989, 86, 1603–1607. [Google Scholar] [CrossRef] [Green Version]
  171. Hu, X.; Fu, M.; Zhao, X.; Wang, W. The JAK/STAT signaling pathway: From bench to clinic. Signal Transduct. Target. Ther. 2021, 6, 402. [Google Scholar] [CrossRef]
  172. Howell, M.D.; Kuo, F.I.; Smith, P.A. Targeting the Janus kinase family in autoimmune skin diseases. Front. Immunol. 2019, 10, 2342. [Google Scholar] [CrossRef]
  173. Shiratori-Hayashi, M.; Yamaguchi, C.; Eguchi, K.; Shiraishi, Y.; Kohno, K.; Mikoshiba, K.; Inoue, K.; Nishida, M.; Tsuda, M. Astrocytic STAT3 activation and chronic itch require IP3R1/TRPC-dependent Ca2+ signals in mice. J. Allergy Clin. Immunol. 2021, 147, 1341–1353. [Google Scholar] [CrossRef]
  174. Oetjen, L.K.; Mack, M.R.; Feng, J.; Whelan, T.M.; Niu, H.; Guo, C.J.; Chen, S.; Trier, A.M.; Xu, A.Z.; Tripathi, S.V. Sensory neurons co-opt classical immune signaling pathways to mediate chronic itch. Cell 2017, 171, 217–228.e213. [Google Scholar] [CrossRef] [Green Version]
  175. Kim, B.S. The translational revolution of itch. Neuron 2022, in press. [Google Scholar] [CrossRef]
  176. Tsuda, M.; Kohro, Y.; Yano, T.; Tsujikawa, T.; Kitano, J.; Tozaki-Saitoh, H.; Koyanagi, S.; Ohdo, S.; Ji, R.-R.; Salter, M.W. JAK-STAT3 pathway regulates spinal astrocyte proliferation and neuropathic pain maintenance in rats. Brain 2011, 134, 1127–1139. [Google Scholar] [CrossRef]
  177. Dominguez, E.; Rivat, C.; Pommier, B.; Mauborgne, A.; Pohl, M. JAK/STAT3 pathway is activated in spinal cord microglia after peripheral nerve injury and contributes to neuropathic pain development in rat. J. Neurochem. 2008, 107, 50–60. [Google Scholar] [CrossRef]
  178. Crispino, N.; Ciccia, F. JAK/STAT pathway and nociceptive cytokine signalling in rheumatoid arthritis and psoriatic arthritis. Clin. Exp. Rheumatol. 2021, 39, 668–675. [Google Scholar] [CrossRef]
  179. Salaffi, F.; Carotti, M.; Farah, S.; Ceccarelli, L.; Giovagnoni, A.; Di Carlo, M. Early response to JAK inhibitors on central sensitization and pain catastrophizing in patients with active rheumatoid arthritis. Inflammopharmacology 2022, 30, 1119–1128. [Google Scholar] [CrossRef]
  180. Reichardt, L.F. Neurotrophin-regulated signalling pathways. Philos. Trans. R. Soc. B Biol. Sci. 2006, 361, 1545–1564. [Google Scholar] [CrossRef] [Green Version]
  181. Sofroniew, M.V.; Howe, C.L.; Mobley, W.C. Nerve growth factor signaling, neuroprotection, and neural repair. Annu. Rev. Neurosci. 2001, 24, 1217–1281. [Google Scholar] [CrossRef]
  182. Aloe, L.; Rocco, M.L.; Balzamino, B.O.; Micera, A. Nerve growth factor: A focus on neuroscience and therapy. Curr. Neuropharmacol. 2015, 13, 294–303. [Google Scholar] [CrossRef] [Green Version]
  183. Aloe, L.; Rocco, M.L.; Bianchi, P.; Manni, L. Nerve growth factor: From the early discoveries to the potential clinical use. J. Transl. Med. 2012, 10, 239. [Google Scholar] [CrossRef] [Green Version]
  184. Micera, A.; Puxeddu, I.; Aloe, L.; Levi-Schaffer, F. New insights on the involvement of Nerve Growth Factor in allergic inflammation and fibrosis. Cytokine Growth Factor Rev. 2003, 14, 369–374. [Google Scholar] [CrossRef]
  185. Lambiase, A.; Micera, A.; Sgrulletta, R.; Bonini, S.; Bonini, S. Nerve growth factor and the immune system: Old and new concepts in the cross-talk between immune and resident cells during pathophysiological conditions. Curr. Opin. Allergy Clin. Immunol. 2004, 4, 425–430. [Google Scholar] [CrossRef]
  186. Yamaguchi, J.; Aihara, M.; Kobayashi, Y.; Kambara, T.; Ikezawa, Z. Quantitative analysis of nerve growth factor (NGF) in the atopic dermatitis and psoriasis horny layer and effect of treatment on NGF in atopic dermatitis. J. Dermatol. Sci. 2009, 53, 48–54. [Google Scholar] [CrossRef]
  187. Webb, M.P.; Helander, E.M.; Menard, B.L.; Urman, R.D.; Kaye, A.D. Tanezumab: A selective humanized mAb for chronic lower back pain. Ther. Clin. Risk Manag. 2018, 14, 361. [Google Scholar] [CrossRef] [Green Version]
  188. Dakin, P.; Kivitz, A.J.; Gimbel, J.S.; Skrepnik, N.; DiMartino, S.J.; Emeremni, C.A.; Gao, H.; Stahl, N.; Weinreich, D.M.; Yancopoulos, G.D. Efficacy and safety of fasinumab in patients with chronic low back pain: A phase II/III randomised clinical trial. Ann. Rheum. Dis. 2021, 80, 509–517. [Google Scholar] [CrossRef]
  189. Déry, O.; Corvera, C.U.; Steinhoff, M.; Bunnett, N.W. Proteinase-activated receptors: Novel mechanisms of signaling by serine proteases. Am. J. Physiol. Cell Physiol. 1998, 274, C1429–C1452. [Google Scholar] [CrossRef]
  190. Steinhoff, M.; Vergnolle, N.; Young, S.; Tognetto, M.; Amadesi, S.; Ennes, H.; Trevisani, M.; Hollenberg, M.; Wallace, J.; Caughey, G. Agonists of proteinase-activated receptor 2 induce inflammation by a neurogenic mechanism. Nat. Med. 2000, 6, 151–158. [Google Scholar] [CrossRef]
  191. Rothmeier, A.S.; Ruf, W. Protease-activated receptor 2 signaling in inflammation. Proc. Semin. Immunopathol. 2012, 34, 133–149. [Google Scholar] [CrossRef]
  192. Shpacovitch, V.; Feld, M.; Bunnett, N.; Steinhoff, M. Protease-activated receptors: Novel PARtners in innate immunity. Trends Immunol. 2007, 28, 541–550. [Google Scholar] [CrossRef]
  193. Liu, Q.; Weng, H.-J.; Patel, K.N.; Tang, Z.; Bai, H.; Steinhoff, M.; Dong, X. The distinct roles of two GPCRs, MrgprC11 and PAR2, in itch and hyperalgesia. Sci. Signal. 2011, 4, ra45. [Google Scholar] [CrossRef] [Green Version]
  194. Zhao, J.; Munanairi, A.; Liu, X.-Y.; Zhang, J.; Hu, L.; Hu, M.; Bu, D.; Liu, L.; Xie, Z.; Kim, B.S. PAR2 mediates itch via TRPV3 signaling in keratinocytes. J. Investig. Dermatol. 2020, 140, 1524–1532. [Google Scholar] [CrossRef]
  195. Yau, M.-K.; Liu, L.; Fairlie, D.P. Toward drugs for protease-activated receptor 2 (PAR2). J. Med. Chem. 2013, 56, 7477–7497. [Google Scholar] [CrossRef]
  196. Fiorucci, S.; Distrutti, E. Role of PAR2 in pain and inflammation. Trends Pharmacol. Sci. 2002, 23, 153–155. [Google Scholar] [CrossRef]
  197. Vergnolle, N.; Bunnett, N.; Sharkey, K.; Brussee, V.; Compton, S.; Grady, E.; Cirino, G.; Gerard, N.; Basbaum, A.; Andrade-Gordon, P. Proteinase-activated receptor-2 and hyperalgesia: A novel pain pathway. Nat. Med. 2001, 7, 821–826. [Google Scholar] [CrossRef]
  198. Lam, D.; Schmidt, B. Serine proteases and protease-activated receptor 2-dependent allodynia: A novel cancer pain pathway. PAIN 2010, 149, 263–272. [Google Scholar] [CrossRef] [Green Version]
  199. Buhl, T.; Ikoma, A.; Kempkes, C.; Cevikbas, F.; Sulk, M.; Buddenkotte, J.; Akiyama, T.; Crumrine, D.; Camerer, E.; Carstens, E. Protease-activated receptor-2 regulates neuro-epidermal communication in atopic dermatitis. Front. Immunol. 2020, 11, 1740. [Google Scholar] [CrossRef]
  200. Lam, D.K.; Dang, D.; Zhang, J.; Dolan, J.C.; Schmidt, B.L. Novel animal models of acute and chronic cancer pain: A pivotal role for PAR2. J. Neurosci. 2012, 32, 14178–14183. [Google Scholar] [CrossRef]
  201. McIntosh, K.A.; Cunningham, M.R.; Bushell, T.; Plevin, R. The development of proteinase-activated receptor-2 modulators and the challenges involved. Biochem. Soc. Trans. 2020, 48, 2525–2537. [Google Scholar] [CrossRef]
  202. Wei, H.; Wei, Y.; Tian, F.; Niu, T.; Yi, G. Blocking proteinase-activated receptor 2 alleviated neuropathic pain evoked by spinal cord injury. Physiol. Res. 2016, 65, 145. [Google Scholar] [CrossRef]
  203. Andoh, T.; Takayama, Y.; Yamakoshi, T.; Lee, J.-B.; Sano, A.; Shimizu, T.; Kuraishi, Y. Involvement of serine protease and proteinase-activated receptor 2 in dermatophyte-associated itch in mice. J. Pharmacol. Exp. Ther. 2012, 343, 91–96. [Google Scholar] [CrossRef] [Green Version]
  204. Egeo, G.; Fofi, L.; Barbanti, P. Botulinum neurotoxin for the treatment of neuropathic pain. Front. Neurol. 2020, 11, 716. [Google Scholar] [CrossRef]
  205. Hary, V.; Schitter, S.; Martinez, V. Efficacy and safety of botulinum A toxin for the treatment of chronic peripheral neuropathic pain: A systematic review of randomized controlled trials and meta-analysis. Eur. J. Pain 2022, 26, 980–990. [Google Scholar] [CrossRef]
  206. Portugal, D.M.; Ferreira, E.F.; Camões-Barbosa, A. Botulinum toxin type A therapy for bilateral focal neuropathic pruritus in multiple sclerosis: A case report. Int. J. Rehabil. Res. 2021, 44, 382–383. [Google Scholar] [CrossRef]
  207. Maari, C.; Marchessault, P.; Bissonnette, R. Treatment of notalgia paresthetica with botulinum toxin A: A double-blind randomized controlled trial. J. Am. Acad. Dermatol. 2014, 70, 1139–1141. [Google Scholar] [CrossRef]
  208. Shaarawy, E.; Hegazy, R.A.; Abdel Hay, R.M. Intralesional botulinum toxin type A equally effective and better tolerated than intralesional steroid in the treatment of keloids: A randomized controlled trial. J. Cosmet. Dermatol. 2015, 14, 161–166. [Google Scholar] [CrossRef]
  209. Klager, S.; Kumar, M.G. Treatment of pruritus with botulinum toxin in a pediatric patient with Fox-Fordyce disease. Pediatric Dermatol. 2021, 38, 950–951. [Google Scholar] [CrossRef]
  210. Meixiong, J.; Dong, X.; Weng, H.-J. Neuropathic itch. Cells 2020, 9, 2263. [Google Scholar] [CrossRef]
  211. Boozalis, E.; Sheu, M.; Selph, J.; Kwatra, S.G. Botulinum toxin type A for the treatment of localized recalcitrant chronic pruritus. J. Am. Acad. Dermatol. 2018, 78, 192–194. [Google Scholar] [CrossRef] [Green Version]
  212. Villamil, A.G.; Bandi, J.C.; Galdame, O.A.; Gerona, S.; Gadano, A.C. Efficacy of lidocaine in the treatment of pruritus in patients with chronic cholestatic liver diseases. Am. J. Med. 2005, 118, 1160–1163. [Google Scholar] [CrossRef]
  213. Lee, H.G.; Grossman, S.K.; Valdes-Rodriguez, R.; Berenato, F.; Korbutov, J.; Chan, Y.-H.; Lavery, M.J.; Yosipovitch, G. Topical ketamine-amitriptyline-lidocaine for chronic pruritus: A retrospective study assessing efficacy and tolerability. J. Am. Acad. Dermatol. 2017, 76, 760–761. [Google Scholar] [CrossRef] [Green Version]
  214. Kopecky, E.A.; Jacobson, S.; Hubley, P.; Palozzi, L.; Clarke, H.M.; Koren, G. Safety and pharmacokinetics of EMLA in the treatment of postburn pruritus in pediatric patients: A pilot study. J. Burn Care Rehabil. 2001, 22, 235–242. [Google Scholar] [CrossRef]
  215. Young, T.A.; Patel, T.S.; Camacho, F.; Clark, A.; Freedman, B.I.; Kaur, M.; Fountain, J.; Williams, L.L.; Yosipovitch, G.; Fleischer, A.B., Jr. A pramoxine-based anti-itch lotion is more effective than a control lotion for the treatment of uremic pruritus in adult hemodialysis patients. J. Dermatol. Treat. 2009, 20, 76–81. [Google Scholar] [CrossRef]
  216. Sutton, K.; Martin, D.; Pinnock, R.; Lee, K.; Scott, R. Gabapentin inhibits high-threshold calcium channel currents in cultured rat dorsal root ganglion neurones. Br. J. Pharmacol. 2002, 135, 257–265. [Google Scholar] [CrossRef] [Green Version]
  217. Matsuda, K.M.; Sharma, D.; Schonfeld, A.R.; Kwatra, S.G. Gabapentin and pregabalin for the treatment of chronic pruritus. J. Am. Acad. Dermajtol. 2016, 75, 619–625.e616. [Google Scholar] [CrossRef]
Figure 1. Interaction between pain and itch transmission in physiological conditions. The activation of itch primary afferents in dorsal root ganglia by pruritogens stimulates the release of excitatory neurotransmitters from the terminals of the secondary itch neurons in the spinal cord, leading to the release of GRP and opioids to activate GRPR+ interneurons for itch transmission. Stimulation of nociceptors induces the activation of secondary nociceptive neurons in the spinal cord for nociceptive transduction. Simultaneously, the activation of nociceptors results in the subsequent activation of Bhlhb5+ inhibitory neurons to compress itch transmission in GRPR+ neurons. Furthermore, spinal cord opioids can activate k-opioid receptors to suppress both pain and itch via reducing µ-opioid receptor activity and enhancing the activity of Bhlhb5+ inhibitory neurons. GRPR—gastrin-related peptide receptor; GRP—gastrin-related peptide; Bhlhb5—Class B basic helix-loop-helix protein 5.
Figure 1. Interaction between pain and itch transmission in physiological conditions. The activation of itch primary afferents in dorsal root ganglia by pruritogens stimulates the release of excitatory neurotransmitters from the terminals of the secondary itch neurons in the spinal cord, leading to the release of GRP and opioids to activate GRPR+ interneurons for itch transmission. Stimulation of nociceptors induces the activation of secondary nociceptive neurons in the spinal cord for nociceptive transduction. Simultaneously, the activation of nociceptors results in the subsequent activation of Bhlhb5+ inhibitory neurons to compress itch transmission in GRPR+ neurons. Furthermore, spinal cord opioids can activate k-opioid receptors to suppress both pain and itch via reducing µ-opioid receptor activity and enhancing the activity of Bhlhb5+ inhibitory neurons. GRPR—gastrin-related peptide receptor; GRP—gastrin-related peptide; Bhlhb5—Class B basic helix-loop-helix protein 5.
Pharmaceuticals 15 00892 g001
Figure 2. Interaction between pain and itch signaling in pathological (chronic) conditions relating to druggable targets with both antinociceptive and antipruritic effects. Multiple elements can contribute to the development of the mismatch and sensitization of chronic pain and itch in pathological conditions. (1) Afferents involved in pain and itch can be activated by either pruritogens or algogens, leading to the mismatched activation of pruriceptors or nociceptors, and partly account for peripheral sensitization. Moreover, central sensitization of spinal cord neurons is associated with (2) the upregulation of GRP and GRPR, and (3) reduction or loss of inhibitory control from Bhlhb5+ inhibitory neurons. Subsequently, these events disrupt the normal interaction between itch and pain. Arrow—activation; diamond—inhibition.
Figure 2. Interaction between pain and itch signaling in pathological (chronic) conditions relating to druggable targets with both antinociceptive and antipruritic effects. Multiple elements can contribute to the development of the mismatch and sensitization of chronic pain and itch in pathological conditions. (1) Afferents involved in pain and itch can be activated by either pruritogens or algogens, leading to the mismatched activation of pruriceptors or nociceptors, and partly account for peripheral sensitization. Moreover, central sensitization of spinal cord neurons is associated with (2) the upregulation of GRP and GRPR, and (3) reduction or loss of inhibitory control from Bhlhb5+ inhibitory neurons. Subsequently, these events disrupt the normal interaction between itch and pain. Arrow—activation; diamond—inhibition.
Pharmaceuticals 15 00892 g002
Table 1. Targets and therapeutic compounds with antinociceptive and antipruritic effects. Abbreviations: KOR—k-opioid receptor; MOR—µ-opioid receptor; H1R—histamine H1 receptor; H4R—histamine H4 receptor; CB1—cannabinoid receptor type 1; CB2—cannabinoid receptor type 2; NGF—nerve growth factor; PAR2—protease-activated receptor 2.
Table 1. Targets and therapeutic compounds with antinociceptive and antipruritic effects. Abbreviations: KOR—k-opioid receptor; MOR—µ-opioid receptor; H1R—histamine H1 receptor; H4R—histamine H4 receptor; CB1—cannabinoid receptor type 1; CB2—cannabinoid receptor type 2; NGF—nerve growth factor; PAR2—protease-activated receptor 2.
TargetEffects on Itch and PainTherapeutic Compounds
TRPV1↑ pain; ↑ itchAsivatrep, AG1529, NGX-4010
TRPV3↑ pain; ↑ itchCitrusinine-II, dyclonine, FTP-THQ
TRPV4↑ pain; ↑ itchisopropyl cyclohexane
TRPA1↑ pain; ↑ itch
TRPM8↑↓ pain; ↑ itchDi-isopropyl-phosphinoyl-alkanes, WS-12
TRPC3↑ pain; ↑ itch
KOR ↓ pain; ↓ itchDifelikefalin, HSK21542, triazole 1.1, butorphanol, nalbuphine
MOR↓ pain; ↑ itchBuprenorphinenaloxone, nalbuphine
H1R↑ pain; ↑ itchChlorpheniramine, fexofenadine, promethazine, diphenhydramine, orphenadrine, mepyramine, pyrilamine
H4R↓ pain(central); ↑ pain(peripheral); ↑ itchJNJ7777120
CB1↓ pain; ↓ itchDronabinol
CB2↓ pain; ↓ itchDronabinol
Oncostatin M↑ pain; ↑ itch
JAK-STAT signaling↑ pain; ↑ itchBaricitinib, upadacitinib
NGF↑ pain; ↑ itch
PAR2↑ pain; ↑ itchFSLLRY-NH2
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Weng, H.-J.; Pham, Q.T.T.; Chang, C.-W.; Tsai, T.-F. Druggable Targets and Compounds with Both Antinociceptive and Antipruritic Effects. Pharmaceuticals 2022, 15, 892. https://doi.org/10.3390/ph15070892

AMA Style

Weng H-J, Pham QTT, Chang C-W, Tsai T-F. Druggable Targets and Compounds with Both Antinociceptive and Antipruritic Effects. Pharmaceuticals. 2022; 15(7):892. https://doi.org/10.3390/ph15070892

Chicago/Turabian Style

Weng, Hao-Jui, Quoc Thao Trang Pham, Chia-Wei Chang, and Tsen-Fang Tsai. 2022. "Druggable Targets and Compounds with Both Antinociceptive and Antipruritic Effects" Pharmaceuticals 15, no. 7: 892. https://doi.org/10.3390/ph15070892

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