Skip to main content
British Journal of Pharmacology logoLink to British Journal of Pharmacology
. 2012 May;166(2):510–521. doi: 10.1111/j.1476-5381.2012.01851.x

The functions of TRPA1 and TRPV1: moving away from sensory nerves

ES Fernandes 1, MA Fernandes 2, JE Keeble 2
PMCID: PMC3417484  PMID: 22233379

Abstract

The transient receptor potential vanilloid 1 and ankyrin 1 (TRPV1 and TRPA1, respectively) channels are members of the TRP superfamily of structurally related, non-selective cation channels. It is rapidly becoming clear that the functions of TRPV1 and TRPA1 interlink with each other to a considerable extent. This is especially clear in relation to pain and neurogenic inflammation where TRPV1 is coexpressed on the vast majority of TRPA1-expressing sensory nerves and both integrate a variety of noxious stimuli. The more recent discovery that both TRPV1 and TRPA1 are expressed on a multitude of non-neuronal sites has led to a plethora of research into possible functions of these receptors. Non-neuronal cells on which TRPV1 and TRPA1 are expressed vary from vascular smooth muscle to keratinocytes and endothelium. This review will discuss the expression, functionality and roles of these non-neuronal TRP channels away from sensory nerves to demonstrate the diverse nature of TRPV1 and TRPA1 in addition to a direct role in pain and neurogenic inflammation.

Keywords: TRPV1, TRPA1, capsaicin, non-neuronal, expression, functionality

Introduction

The transient receptor potential vanilloid 1 and ankyrin 1 (TRPV1 and TRPA1, respectively) channels are members of the TRP superfamily of structurally related, non-selective cation channels that is divided into seven families: TRPC (Canonical), TRPN (no mechanoreceptor potential C), TRPM (Melastatin), TRPML (Mucolipin), TRPP (Polycystin), TRPA (Ankyrin) and TRPV (Vanilloid), each present in several species across the animal kingdom (Pedersen et al., 2005; Clapham, 2007). TRP channels have many different physiological roles, ranging from purported roles as store-operated calcium channels (e.g. TRPC3, TRPC7; Riccio et al., 2002; Kaznacheyeva et al., 2007), to roles in thermo- (e.g. A1, M8, V1, V4; Caterina et al., 1997; Güler et al., 2002; Peier et al., 2002; Story et al., 2003), mechano- (e.g. A1, C1, V1, V4; Liedtke et al., 2003; Walker et al., 2003; Corey et al., 2004; Maroto et al., 2005) and chemo-sensation (e.g. A1, M8, V1; Peier et al., 2002; Bandell et al., 2004; Andersson et al., 2008). All TRP channels are tetramers formed by subunits with six transmembrane domains and cation-selective pores, which frequently show high calcium permeability (Latorre et al., 2009).

Whilst all seven families have wide roles in many physiological and pathophysiological processes, TRPV1 and TRPA1 will be the specific focus of this review. Both TRPV1 and TRPA1 play an integral role in pain (Bevan and Andersson, 2009; Fernandes et al., 2011) and neurogenic inflammation (Geppetti et al., 2008) via sensory nerve activation. In fact, 97% of TRPA1-expressing sensory neurons express TRPV1, while 30% of TRPV1-expressing neurons express TRPA1 (Story et al., 2003). Both TRPV1 and TRPA1 channels are calcium-permeable, and, although their subunits have not been shown to form a heterotetramer channel, they may form a complex on the plasma membrane of sensory neurons. This enables TRPV1 to influence intrinsic characteristics of the TRPA1 channel, including voltage–current relationships and its open probability at negative holding potentials (Staruschenko et al., 2010). Similarly, Salas et al. (2009) have shown that the features of neuronal TRPA1 are not duplicated in cells expressing only TRPA1 and, instead, can be restored only when TRPA1 and TRPV1 channels are coexpressed. Moreover, both TRPV1 and TRPA1 are integrators of a range of noxious stimuli, and TRPV1 and TRPA1 agonists are able, at least in part, to heterologously desensitize TRPV1 and TRPA1 pathways (Ruparel et al., 2008). Overall, based on evidence such as that described above, it may well be that TRPV1 and TRPA1 are ‘partners in crime’ in the activation of sensory nerves.

The role of these TRP channels in pain and neurogenic inflammation have, to date, been very well covered by previous authors (e.g. Bevan and Andersson, 2009; Cortright and Szallasi, 2009; Stucky et al., 2009; Fernandes et al., 2011), which reflects the enormity of the role that these channels play in sensory nerve function at both a central and peripheral level. However, there is accumulating evidence that TRPV1 and TRPA1 have functional roles away from sensory nerve activity, which this review aims to address. In order to substantiate a role for these non-sensory nerve TRP channels in either normal physiology or disease, several criteria must be fulfilled: (1) TRPV1/TRPA1 expression should be demonstrated in these tissues at the gene and protein level; (2) the receptors should be functional, that is, showing evidence of calcium permeability; and (3) a role in either physiological or pathophysiological functions should be demonstrated. In discussing these other functions, the imperative importance of TRPV1 and TRPA1 in pain and neurogenic inflammation cannot and should not be diminished. Furthermore, it is entirely possible that the TRPV1 and TRPA1 activity away from sensory nerves may indirectly affect pain and neurogenic inflammation.

TRPV1

The vanilloid TRP channels are divided into six members – 1 to 6. However, only TRPV1 from the TRPV subfamily is actually activated by vanilloids, including capsaicin, the pungent component of chilli peppers. The capacity of capsaicin to activate sensory nerves was determined by Jancsó in the 1960s (Jancsóet al., 1967). However, it was 15 years later that the presence of a capsaicin receptor in the plasma membrane of sensory nerves was recognized by Szolcsányi and Jancsó-Gábor (1975). In 1997 came perhaps the most significant advance in TRPV1 research, when the mouse TRPV1 receptor was cloned by Caterina et al. (1997) [later cloned in humans (Hayes et al., 2000) and guinea-pigs (Savidge et al., 2002)]. With this major advance in the understanding of TRPV1 at a molecular level came the confirmation that heat >43°C and pH <5.9 also activate TRPV1 (Tominaga et al., 1998). Furthermore, the generation of TRPV1 knockout mice and a comprehensive analysis of the phenotype of the animals confirmed a pivotal role for this receptor in noxious heat sensation in vivo (Caterina et al., 2000). A variety of exogenous and endogenous activators of TRPV1 have since been identified (see Table 1).

Table 1.

Agonists and modulators

TRPV1 TRPA1
Exogenous agonists Capsaicin (Caterina et al., 1997) Allyl Isothiocyanate (Bandell et al., 2004)
Vanilloids (e.g. Olvanil, Resiniferatoxin) (Brand et al., 1987; Szallasi and Blumberg, 1989) Environmental pollutants, e.g. acrolein (Bautista et al., 2006)
Capsinoids (e.g. Capsiate) (Ohnuki et al., 2001) Irritants (e.g. Formalin) (Mcnamara et al., 2007)
Noxious high temperature (>43°C) (Caterina et al., 1997) Cold temperatures (<17°C) (Kwan et al., 2006; Karashima et al., 2009)
Low pH (<6.0) (Caterina et al., 1997) Allicin (Macpherson et al., 2005)
Camphor (Xu et al., 2005) Icilin (Story et al., 2003)
Cinnamaldehyde (Macpherson et al., 2006)
Tetrahydrocannabinol (Jordt et al., 2004)
Endogenous agonists Anandamide (Zygmunt et al., 1999; Smart et al., 2000) Oxidative stress products (Bessac et al., 2008) e.g. 4-hydroxynonenal (Trevisani et al., 2007).
Lipoxygenase products (e.g. LTB4) (Hwang et al., 2000; Huang et al., 2002) Lipid peroxidation products (Taylor-Clark et al., 2008)
N-acyldopamines (Huang et al., 2002; Chu et al., 2003) Zinc, copper and cadmium (Hu et al., 2009; Gu and Lin, 2010)
Endogenous modulators (via activation of intracellular pathways) Bradykinin (Chuang et al., 2001) Bradykinin (Bandell et al., 2004; S. Wang et al., 2008)
PAR-2 agonists (Amadesi et al., 2004) PAR-2 agonists (Dai et al., 2007)
NGF (Chuang et al., 2001)
ATP (Chuang et al., 2001)

In this context, it is important to mention that the TRPV1 receptor undergoes desensitisation after repeated administration of, or prolonged exposure to, capsaicin (Szolcsányi et al., 1975) or resinaferatoxin (Szolcsányi et al., 1990). The receptor not only becomes desensitized to other activators of the same receptor, but the TRPV1 pathway heterologously desensitizes responses to TRPA1 agonists and vice versa (Ruparel et al., 2008). The mechanisms underlying homologous and heterologous desensitization appear to be distinct with capsaicin-induced desensitization being calcium-dependent but heterologous desensitization being calcium-independent (Ruparel et al., 2008). In addition, high doses of capsaicin selectively destroy C- and Aδ sensory nerves, thereby completely preventing nerve activation (Szallasi et al., 1995). Clearly, this means that capsaicin-induced desensitization/nerve destruction cannot be used to examine TRPV1-specific down-stream effects on physiological/pathophysiological processes. However, it can be used to show that capsaicin-induced desensitization via TRPV1 has a functional effect upon a cell or tissue.

TRPA1

Although TRPA1 was identified slightly later than its TRPV1 counterpart, its discovery took a comparable course. The effects of the classic TRPA1 agonist, mustard oil (the active component being allyl isothiocyanate), were identified much earlier (e.g. Koltzenburg and McMahon, 1986) than the receptor itself, with this compound commonly being used to induce neurogenic inflammation, in a similar manner to capsaicin. TRPA1 was first cloned by Jaquemar et al. (1999), although its expression on neurons was not reported until 2003 when Story et al. identified it as a receptor for noxious cold temperature. However, the association between TRPA1 and mustard oil was later confirmed in 2005 when Jordt and colleagues not only established TRPA1 as the mustard oil receptor but also for tetrahydrocannabinol, leading the way into a plethora of research into further possible TRPA1 agonists (see Table 1). Indeed, concomitant with the groundbreaking generation of TRPA1 knockout mice (Bautista et al., 2006) came the finding that TRPA1 mediates the effects of acrolein (2-propenal), present in tear gas, vehicle exhaust, tobacco products and byproducts of chemotherapeutic agents. Unlike the situation with TRPV1 knockout mice, an examination of the phenotype of TRPA1 knockouts regarding a role for this receptor in noxious temperature sensation has proved controversial. Two independent knockout animals have been generated (Bautista et al., 2006; Kwan et al., 2006). The portion of the gene knocked out was the fifth and sixth transmembrane domains (required for ion conduction) in both cases. However, Bautista et al. (2006) reported normal responses to cold temperatures both in vivo and in vitro in TRPA1 knockout mice. On the other hand, knockouts in the Kwan et al. (2006) study displayed a clear functional deficit to cold stimulation. A discussion of the reasons underlying these differences is beyond the scope of this review, but it remains to be seen whether TRPA1 is truly a thermosensor channel.

TRPV1 and TRPA1 expression

Central expression

The location of TRPV1 on small-diameter, Aδ and C fibre sensory nerves is clearly intrinsically linked with its role in pain and neurogenic inflammation. These nerves were the site of discovery for both TRPV1 and TRPA1 and are still today the focus of the bulk of research on these TRP channels. With sensory nerves, the effects of TRPV1 (and to a certain extent TRPA1) activation were known well before the receptor or its expression was identified. Effects of capsaicin in other cell types were also postulated as early as the 1970s (e.g. Jancsô and Wollemann, 1977). However, since the cloning of the receptors and the revolution of molecular biology, TRPV1 and TRPA1 expression on other cell types has been confirmed as an essential basis of the pharmacology of these effects. For example, Mezey et al. (2000) used a TRPV1-specific antibody to demonstrate the presence of TRPV1-expressing neurons throughout the neuroaxis, including such areas as the dopaminergic neurones of the substantia nigra, hippocampal pyramidal neurones, hypothalamic neurones and neurones in the locus coeruleus, in addition to various layers of the cortex. RT-PCR confirmed the expression of TRPV1 mRNA in the hippocampus, hypothalamus and cortex (Mezey et al., 2000). The presence of TRPV1 mRNA has also been identified in the cerebellum (Sasamura et al., 1998), showing a widespread of expression of TRPV1 within the CNS. However, more recently, the use of TRPV1 reporter mice has revolutionized the study of TRPV1 expression, and they would suggest that the expression of this receptor is minimal within a few discrete brain regions, most obviously in the vicinity of the caudal hypothalamus (Cavanaugh et al., 2011).

The central localization of TRPA1 has been less specific than for TRPV1, although TRPA1 mRNA has been shown to be abundant in dog brain and cerebellum (Doihara et al., 2009). However, associated protein expression was not confirmed in this study.

Peripheral expression

Within the periphery, recent evidence has located TRPV1 and TRPA1 on a variety of non-neuronal tissues. In fact, the list of possible sites of expression is becoming so great that to discuss all in turn would be outside of the scope of this article. Tables 2 and 3 list the peripheral non-neuronal cells on which TRPV1 and TRPA1 have been located by RT-PCR for RNA expression and/or immunohistochemical staining Western blot for protein expression. From this extensive list, it is clear to see that the advent of molecular biology techniques in this field a decade ago has revolutionized our potential understanding of the possible functions of these TRP channels. However, as mentioned previously, evidence that receptor mRNA or protein is present in a tissue should be substantiated by evidence that the channel is functional. Therefore, the following section will address that studies that have provided evidence of TRPV1/TRPA1 functionality in brain or peripheral sites.

Table 2.

TRPV1 expression

Expression site Immunostaining/immunofluorescence) RT-PCR Western blot [Ca2+] functionality Possible role/effects on activation References
TRPV1
Mouse
 Arteriolar smooth muscle cells Vasoconstriction Cavanaugh et al., 2011
 Mesenteric arteries and endothelial cells Vasorelaxation Yang et al., 2010
 Laryngeal epithelium Laryngeal nociceptors Hamamoto et al., 2008 (mouse); Hamamoto et al., 2009 (human)
 Preadipocytes and adipose tissue Adipogenesis Zhang et al., 2007b
 Urothelium Stretch-evoked ATP release Birder et al., 2001
Rat
 Vascular smooth muscle Vasoconstriction Kark et al., 2008
 Pulmonary artery smooth muscle Vasoconstriction Yang et al., 2006
 Pancreatic B cells Increased insulin secretion Akiba et al., 2004
Human
 Corneal epithelium Inflammatory mediator secretion Zhang et al., 2007a
 Corneal endothelium Temperature sensation Mergler et al., 2010
 Cerebromicrovascular endothelium Regulation of blood brain barrier permeability Golech et al., 2004
 Blood Nociception; role in inflammatory processes? Saunders et al., 2007
 Mononuclear cells
 Epidermal keratinocytes Noxious chemical sensor Inoue et al., 2002
 Preadipocytes and adipose tissue Adipogenesis Zhang et al., 2007b
 Synoviocytes Adaptive/pathological changes in arthritic inflammation Kochukov et al., 2006
 Nasal vascular endothelium, epithelials and submucosal glands Stimulate epithelial secretions Seki et al., 2006

Table 3.

TRPA1 expression

Expression site Immunostaining/immunofluorescence) RT-PCR Western blot [Ca2+] functionality Possible role/effects on activation References
TRPA1
Mouse
 Auditory hair cell; organ of corti; utricle, saccule and crista ampullaris Mechanosensor? Nagata et al., 2005
 Enterochromaffin cells Regulates gastrointestinal motility via 5HT release Nozawa et al., 2009
 Hair follicle keratinocytes Modulation of cutaneous nerve firing Kwan et al., 2009
Rat
 Cerebral and cerebellar artery endothelium Vasodilation Earley et al., 2009
 Urothelium Detrusor overactivity Streng et al., 2008
 Enterochromaffin cells Regulates gastrointestinal motility via 5HT release Nozawa et al., 2009
Human
 Undifferentiated keratinocytes Thermosensation?? Tsutsumi et al., 2010
 Skin basal keratinocytes ? Anand et al., 2008
 Keratinocytes in epidermis and dermis of hair follicle Keratinocyte differentiation; inflammation; mechano- and thermosensor Atoyan et al., 2009
 Melanocytes
 Fibroblasts
 Synoviocytes Adaptive/pathological changes in arthritic inflammation Kochukov et al., 2006
 Enterochromaffin cells Regulates gastrointestinal motility via 5HT release Nozawa et al., 2009
Other
 Dog brain and cerebellum ? Doihara et al., 2009

Receptor functionality

Although TRPV1/TRPA1 channel expression has been shown in a wide variety of tissues, evidence of functionality has not yet been demonstrated for all of these. Therefore, this section will solely concentrate on cell types where both a molecular and functional presence has been confirmed. One of the first cell types in which functionality was first identified is epidermal keratinocytes. Inoue et al. (2002) demonstrated that both capsaicin and acidification produced elevations in the intracellular calcium concentration in cultured human epidermal keratinocytes. Furthermore, these increases were inhibited by the TRPV1 antagonist, capsazepine (Inoue et al., 2002). Similarly, treatment of human skin fibroblasts with capsaicin induced significant changes in the membrane current and the intracellular calcium level that were antagonized by capsazepine (Kim et al., 2006). More recently, TRPA1 agonists have also been shown to activate calcium currents in both keratinocytes (cold, allyl isothiocyanate and mustard oil; Atoyan et al., 2009) and fibroblasts (cold, allyl isothiocyanate and cinnemaldehyde; Hu et al., 2010). This trend is carried forth with a multitude of other cells, as summarized in Table 2.

It is therefore clear to this point that TRPV1 and TRPA1 are expressed and functional away from sensory nerves. Nevertheless, this does not mean that the channels have similar sensitivities. However, a comparison of the sensitivity of channels expressed on sensory nerves and other tissues regarding agonist stimulation is extremely difficult to extract from the literature. The only study that to date has directly compared neuronal and non-neuronal TRPV1 responses is that by Kark et al. (2008) in vascular tissue. In the vasculature, capsaicin has biphasic effects: at lower concentrations up to 10 nM, dilations are observed in response to neuronal TRPV1 activation. Conversely, at higher capsaicin concentrations between 0.1 and 1 µM, vasoconstriction is observed in response to non-neuronal TRPV1 activation.

In other cell types, capsaicin increased the intracellular calcium concentration of rat vagal neurons within a capsaicin concentration range of 0.1–10 µM (Marsh et al., 1987) and increased the intracellular calcium concentration of human and murine adipocytes within a capsaicin concentration range of 0.01–1 µM (Zhang et al., 2007). The sensitivity of neuronal and non-neuronal TRPV1 to capsaicin activation in these studies therefore looks similar. However, it is impossible to directly compare studies due to differences in experimental conditions, including species differences (rat vs. mouse/human, respectively, for the Marsh and Zhang studies). Thus, the study by Kark et al. (2008) provides the only evidence to date that there may be differences in sensitivity between neuronal and non-neuronal forms of the receptor. The sensitivity of TRPA1 on sensory nerves compared with TRPA1 on other tissues is also very difficult to discuss, as no study has directly compared the same agonist in each type of tissue across a concentration range. Thus, any difference in sensitivity remains to be seen.

Role in physiological or pathophysiological function

From the previous sections, it is clear that the range of cell types expressing functional TRPV1 and TRPA1 is very diverse. As such, it is not surprising that the associated functions of these receptors are also complex. Physiological or pathophysiological effects of non-neuronal TRPV1 and TRPA1 have been implicated in inflammation, infection and immunity, the cardiovascular system and in conditions such as obesity. Meanwhile, neuronal TRPV1 in the brain may have functions in neurogenesis (Jin et al., 2004) and thermoregulation (Jancsó-Gábor et al., 1970b), amongst others. The following sections will discuss the effects of TRPV1 and TRPA1 in the brain as well as non-neuronal TRPV1 and TRPA1 within the context of different physiological/pathophysiological systems. In doing so, the possible significance of these TRP channels away from sensory nerves will be highlighted.

In the brain

As mentioned previously, TRPV1 is expressed throughout the brain and, moreover, these TRPV1 channels appear to be functional upon agonist stimulation. Indeed, downstream of channel activation, Jancsô and Wollemann (1977) have reported that capsaicin stimulates adenylate cyclase activity in the rat cerebral cortex in vitro. Furthermore, direct injection of capsaicin into the preoptic area of the anterior hypothalamus (Jancsó-Gábor et al., 1970b) or i.c.v. region (Dib, 1982) of the rat brain causes hypothermia, suggesting a role for this channel in thermoregulation. Hypothermia is associated with a fall in rectal and hypothalamic temperature, an increased cutaneous temperature (Dib, 1982) and tail skin vasodilation (Jancsó-Gábor et al., 1970a). Vice versa, rats desensitized by hypothalamic injections of high concentrations of capsaicin lose their ability to thermoregulate against overheating of their bodies and respond with an enhanced hyperthermia to strong sensory stimuli such as repeated pinching of the tail (Jancsó-Gábor et al., 1970a). Similarly, systemic administration of TRPV1 antagonists such as AMG517 (Gavva et al., 2007b), AMG0347 (Steiner et al., 2007) and A-425619 (Gavva et al., 2007a) causes an increase in body temperature within approximately 1 h of treatment. However, antagonist-induced hyperthermia may not be mediated by hypothalamic TRPV1 as peripherally-restricted antagonists still have the capacity to cause an increase in body temperature (Tamayo et al., 2008).

In inflammation, infection and immunity

A physiological/pathophysiological role for non-neuronal TRPV1/TRPA1 is perhaps nowhere more apparent than in the case of inflammation, infection and immunity. However, it is important to note that although the effects of these TRP channels are non-neuronal, it is evident that they may well impact indirectly upon pain and/or neurogenic inflammation. As mentioned previously, keratinocytes functionally express both TRPV1 and TRPA1. These cells play an important role in maintaining the integrity of the immune response in skin as well as stimulating cutaneous inflammation via prostanoid and cytokine release (Gröne et al., 2002). TRPV1 activation by capsaicin causes an increase in COX2 expression in human keratinocytes with a concomitant increase in PGE2 levels in vitro (Southall et al., 2003). An increase in IL-8 is also observed (Southall et al., 2003). Similarly, treatment of human keratinocytes with the TRPA1 agonist, icilin, has been shown to increase the expression of pro-inflammatory ILs (IL-1α and IL-1β; Atoyan et al., 2009) as well altering the expression of genes involved in the control of keratinocyte proliferation, differentiation and cell cycle regulation (Atoyan et al., 2009). Stimulation of inflammatory mediator release by TRPV1/TRPA1 agonists from keratinocytes could well have a significant effect upon sensory nerves that have a high density in skin, especially as PGE2 and IL-1 are known to sensitize and/or activate sensory nerve endings (Schaible and Schmidt, 1988; Binshtok et al., 2008). This is therefore a prime example of how non-neuronal TRPV1/TRPA1 may interact with sensory nerves to affect pain and neurogenic inflammation.

As well as keratinocytes, peripheral blood mononuclear cells (PBMCs) are also directly affected by TRPV1/TRPA1 activation. For example, PBMCs undergo apoptosis when stimulated with capsaicin or resinaferatoxin, an effect that is reversed by the TRPV1 antagonist, AM630 (Saunders et al., 2007). In relation to TRPA1, the cinnamaldehyde derivative, 2′-hydroxycinnamaldehyde has been shown to inhibit nitric oxide release and NF-κB activation in macrophages that have been stimulated with LPS (the cell wall component of Gram-negative bacteria; Lee et al., 2005). Similarly, cinnemaldehyde inhibits IL-1β and TNFα release from human monocytes and macrophages that have been stimulated by LPS (Chao et al., 2008). A concomitant reduction of the release of reactive oxygen species from the macrophages is also observed (Chao et al., 2008). It would therefore appear that the effects of TRPV1 and TRPA1 in terms of their direct effects on inflammatory cells appear to be in part anti-inflammatory.

There are various other cells involved in immunity that respond to TRPV1 and TRPA1 activation including bone marrow-derived dendritic cells, where capsaicin leads to dendritic cell maturation and an increase in antigen presentation (Basu and Srivastava, 2005). Furthermore, with regard to TRPA1, cinnamaldehyde has been shown to cause a dose-dependent suppression of the lymphoproliferation in LPS-treated mouse splenocytes (Koh et al., 1998). The same study also showed that the exposure of thymocytes to cinnamaldehyde accelerated T-cell differentiation from CD4 and CD8 double-positive cells to CD4 or CD8 single-positive cells (Koh et al., 1998).

It is therefore clear that TRPV1 and TRPA1 are expressed, functional and are active within cells relevant to inflammation, infection and immunity. Most of the aforementioned studies have been carried out in vitro and so the precise influences of these effects in an in vivo setting are, as yet, far from clear. However, what it is clear from other studies is that TRPV1 at least plays a paradoxical role in inflammation in vivo, for example, exacerbating inflammation in arthritis and yet in experimentally induced sepsis, TRPV1 null mice demonstrate elevated levels of pathological markers in comparison with wild-type mice (Alawi and Keeble, 2010). It cannot be ruled out at this stage that this is due to differing effects of neuronal and non-neuronal TRPV1 channels.

Role in the vasculature

TRPV1 and TRPA1 have been shown to control vascular responses either by the well-established neurogenic response that is mediated by sensory nerves (Geppetti et al., 2008) or via a direct effect on vascular tissue (Kark et al., 2008; Earley et al., 2009). However, the non-neuronal mechanisms involved in mediating vasodilatation and oedema formation following TRPV1 and TRPA1 activation in vivo are unclear. Both endothelial cells and smooth muscle cells express a variety of membrane ion channels to control Ca2+ influx and membrane potential, including the expression of TRPV1 and TRPA1 channels on endothelial cells (Yao and Garland, 2005; Earley et al., 2009) and TRPV1 expression on vascular smooth muscle cells (Kark et al., 2008; Cavanaugh et al., 2011), as described in Table 1. Thus, there is clearly the potential for non-neuronal TRPV1 and TRPA1 to contribute to vasculature control.

TRPV1 on endothelial cells has been shown to regulate the expression and secretion of endothelial cell-derived CGRP, which affords protective effects on endothelial cells (Luo et al., 2008). Furthermore, CGRP is a potent vasodilator (Brain et al., 1985), and this CGRP may therefore impact upon blood pressure. Indeed, TRPV1 activation on sensory nerves also causes CGRP release, leading to a profound decrease in vascular tone (Zygmunt et al., 1999). On the other hand, TRPV1 expressed on vascular smooth muscle appears to cause vasoconstriction. Kark et al. (2008) have shown that capsaicin triggers transient vasoconstriction in isolated pressurised rat skeletal muscle arterioles, which is not abolished by endothelial cell removal or denervation in vivo, indicating the vasoconstriction was mediated by a direct effect of TRPV1 on vascular smooth muscle. Keeble and Brain (2006) have also demonstrated vasoconstrictor responses to capsaicin, albeit in the mouse synovial membrane. More recently, Cavanaugh et al. (2011) have demonstrated vasoconstriction in response to capsaicin in mouse ear arterioles. Interestingly, as mentioned previously, it has also been suggested that capsaicin has biphasic effects on the vasculature: at lower concentrations, capsaicin (up to 10 nM) evokes vasodilation in skin due to sensory nerve activation, whereas higher concentrations (0.1–1 µM) elicit substantial constrictions in skeletal muscle arterioles due to non-neuronal TRPV1 stimulation (Kark et al., 2008). It is unclear whether this difference is due to receptor sensitivity (as discussed earlier with respect to receptor functionality) or a difference in TRPV1 receptor density in the two tissues. Furthermore, it is not entirely clear whether, in order to achieve vasoconstriction, the vasodilator effect of capsaicin first needs to be counteracted. However, it is possible that highly localised TRPV1 activation by endogenous activators means that vasoconstriction or vasodilation are triggered entirely separately, as opposed to treatment with exogenous capsaicin when all TRPV1 is likely to be affected simultaneously.

TRPV1 may also play a role in vascular responses during chronic hypoxia where up-regulation of the TRPV1 gene and protein is observed (Y.X. Wang et al., 2008). Chronic hypoxia has been shown to enhance the ability of human pulmonary artery smooth muscle cells to proliferate and to increase resting levels of cytosolic calcium and capacitative calcium entry with both effects being inhibited in a dose-dependent manner by the TRPV1 antagonist, capsazepine (Y.X. Wang et al., 2008). These results therefore suggest that TRPV1 on smooth muscle may be a critical pathway or mediator in chronic hypoxia-induced vascular changes.

Research into TRPA1 in the vasculature is still in an early phase although mustard oil has been shown to trigger vasodilatation in rat cerebral arteries via a mechanism that appears to involve TRPA1 expressed on endothelial cells (Earley et al., 2009). Mustard oil-induced vasodilation was not mediated by nitric oxide or prostanoids, rather by calcium-activated potassium channels on endothelial cells and inwardly rectifying potassium channels on arterial myocytes. Furthermore, the responses were inhibited by the TRPA1 antagonist, HC-030031 (Earley et al., 2009).

It is therefore clear that non-neuronal TRPV1 and TRPA1 both have the potential to play a role in the physiology or pathophysiology of the vasculature. TRPV1 in general has been shown to play a role in hypertension (Li and Wang, 2003), cardiac ischaemia (Wang and Wang, 2005) and cardiovascular shock (Akabori et al., 2007). However, the relative contribution of neuronal and non-neuronal TRPV1 to these effects is, as yet, far from clear. In the case of TRPA1, the physiological relevance of TRPA1 on sensory nerves in the vasculature has only just been elucidated (Pozsgai et al., 2010), and so we still have a long way to go with our understanding of this channel.

Obesity and thermogenesis

Obesity is one of the most significant health issues in western society due to the morbidity associated with this condition that is increasing in prevalence. Thus, a significant amount of research has been generated to understand its underlying causes and means of treating/preventing the condition. It is known that obesity is induced by the hypertrophy of adipocytes and the recruitment of new adipocytes from precursor cells. These processes are dependent on the regulation of adipocyte differentiation. The TRPV1 receptor is very interesting in this respect as capsaicin has been shown to inhibit adipocyte differentiation in vitro by activation of AMP-activated protein kinase (Hwang et al., 2005). Furthermore, Hsu and Yen (2007) have shown that treatment of preadipocytes with capsaicin decreases the number of normal adipocytes and increases the number of early apoptotic and late apoptotic cells in a dose-dependent manner. Furthermore, treatment of adipocytes with capsaicin was shown to decrease the quantity of intracellular triglycerides and glycerol-3-phosphate dehydrogenase activity (Hsu and Yen, 2007), both biomarkers of adipogenesis.

As with the previous physiological/pathophysiological conditions discussed, it is not known how significantly non-neuronal TRPV1 receptors contribute to the overall effects of TRPV1 activation. However, the overall effect of TRPV1 modulation in obesity is stark. For example, both animal (Zhang et al., 2007) and human (Ohnuki et al., 2001) data have indicated that the consumption of capsaicin- or non-pungent capsiate-containing foods is correlated with a reduced incidence of obesity. Similarly, oral administration of capsaicin alone also suppresses body fat accumulation in mice (Ohnuki et al., 2001), and dietary capsaicin can reduce obesity-induced insulin resistance and hepatic stenosis in mice fed a high fat diet (Kang et al., 2010). Moreover, TRPV1-mediated changes in thermogenesis may have the potential to impact upon obesity, possibly through changes in expression of thermogenic uncoupling proteins, as seen in response to chronic treatment of rats with capsiates (Masuda et al., 2003).

In recent years, a role for TRPV1 in thermoregulation has also been identified which may, at least in part, be due to changes in thermogenesis (for review, see Romanovsky et al., 2009). For many years, capsaicin has been known to cause a centrally mediated hypothermia in mice (Jancsó-Gábor et al., 1970b). In contrast, its intragastric administration enhances thermogenesis and heat diffusion (Masamoto et al., 2009). Similarly, the jejunal administration of non-pungent capsaicin analogues was shown to increase energy expenditure via direct activation of TRPV1 located on intestinal extrinsic nerves (Kawabata et al., 2009). Interestingly, some TRPV1 antagonists cause hyperthermia, associated with increased thermogenesis (Gavva et al., 2007a) through a peripheral mechanism (Tamayo et al., 2008), whilst TRPV1 gene knock down does not affect body temperature in mice (Tóth et al., 2011). and TRPV1 knockout mice exhibit a normal basal body temperature (Steiner et al., 2007). Although this clearly shows a homeostatic role for TRPV1 in thermoregulation, it is beyond the scope of this review to discuss the mechanism underlying TRPV1 antagonist-induced hyperthermia as there is, to date, no direct evidence that it is mediated by non-neuronal TRPV1. It will be extremely interesting in the future to determine whether the mechanism underlying role of TRPV1 in thermoregulation is intrinsically linked with the aforementioned role for TRPV1 in obesity.

Conclusion

To conclude, it is now clear that the roles of TRPV1 and TRPA1 discussed in this review extend far beyond sensory nerves. Not only are TRPV1 and TRPA1 receptors expressed in other neuronal and non-neuronal tissues, but they also exhibit functionality and they are of potential physiological/pathophysiological relevance. It is clear that we still have a great deal to learn about these receptors away from sensory nerves, especially in relation to their precise function in vivo. We also need to find out a great deal more about their influence upon pain and neurogenic inflammation as it is entirely possible that they are intrinsically related. Progress in this field would be greatly enhanced by selective knockout/knockdown of TRPV1/TRPA1 on sensory nerves and/or other specific cell types. Furthermore, it would be interesting to determine whether TRPV1 antagonists, or the currently available TRPA1 antagonists, have any relative specificity for these TRP channels on different cell types. Finally, it remains to be seen whether different modes of activation of TRPV1 and TRPA1 have potentially differing importance depending on the site of TRP channel expression. We eagerly await these answers.

Acknowledgments

ESF is supported by Arthritis Research UK (grant number 19296). MF and JK are supported by a Capacity Building Award in Integrative Mammalian Biology funded by the BBSRC, BPS, HEFCE, KTN, MRC and SFC.

Glossary

PBMC

peripheral blood mononuclear cell

TRPA1

transient receptor potential cation channel subfamily A member 1

TRPV1

transient receptor potential cation channel subfamily V member 1

Conflict of interest

There is no known conflict of interest.

References

  1. Akabori H, Yamamoto H, Tsuchihashi H, Mori T, Fujino K, Shimizu T, et al. Transient receptor potential vanilloid 1 antagonist, capsazepine, improves survival in a rat hemorrhagic shock model. Ann Surg. 2007;245:964–970. doi: 10.1097/01.sla.0000255577.80800.e1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Akiba Y, Kato S, Katsube K, Nakamura M, Takeuchi K, Ishii H, et al. Transient receptor potential vanilloid subfamily 1 expressed in pancreatic islet beta cells modulates insulin secretion in rats. Biochem Biophys Res Commun. 2004;321:219–225. doi: 10.1016/j.bbrc.2004.06.149. [DOI] [PubMed] [Google Scholar]
  3. Alawi K, Keeble J. The paradoxical role of the transient receptor potential vanilloid 1 receptor in inflammation. Pharmacol Ther. 2010;125:181–195. doi: 10.1016/j.pharmthera.2009.10.005. [DOI] [PubMed] [Google Scholar]
  4. Amadesi S, Nie J, Vergnolle N, Cottrell GS, Grady EF, Trevisani M, et al. Protease-activated receptor 2 sensitizes the capsaicin receptor transient receptor potential vanilloid receptor 1 to induce hyperalgesia. J Neurosci. 2004;24:4300–4312. doi: 10.1523/JNEUROSCI.5679-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Anand U, Otto WR, Facer P, Zebda N, Selmer I, Gunthorpe MJ, et al. TRPA1 receptor localisation in the human peripheral nervous system and functional studies in cultured human and rat sensory neurons. Neurosci Lett. 2008;438:221–227. doi: 10.1016/j.neulet.2008.04.007. [DOI] [PubMed] [Google Scholar]
  6. Andersson DA, Gentry C, Moss S, Bevan S. Transient receptor potential A1 is a sensory receptor for multiple products of oxidative stress. J Neurosci. 2008;28:2485–2494. doi: 10.1523/JNEUROSCI.5369-07.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Atoyan R, Shander D, Botchkareva NV. Non-neuronal expression of transient receptor potential type A1 (TRPA1) in human skin. J Invest Dermatol. 2009;129:2312–2315. doi: 10.1038/jid.2009.58. [DOI] [PubMed] [Google Scholar]
  8. Bandell M, Story GM, Hwang SW, Viswanath V, Eid SR, Petrus MJ, et al. Noxious cold ion channel TRPA1 is activated by pungent compounds and bradykinin. Neuron. 2004;41:849–857. doi: 10.1016/s0896-6273(04)00150-3. [DOI] [PubMed] [Google Scholar]
  9. Basu S, Srivastava P. Immunological role of neuronal receptor vanilloid receptor 1 expressed on dendritic cells. Proc Natl Acad Sci U S A. 2005;102:5120–5125. doi: 10.1073/pnas.0407780102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Bautista DM, Jordt SE, Nikai T, Tsuruda PR, Read AJ, Poblete J, et al. TRPA1 mediates the inflammatory actions of environmental irritants and proalgesic agents. Cell. 2006;124:1269–1282. doi: 10.1016/j.cell.2006.02.023. [DOI] [PubMed] [Google Scholar]
  11. Bessac BF, Sivula M, von Hehn CA, Escalera J, Cohn L, Jordt SE. TRPA1 is a major oxidant sensor in murine airway sensory neurons. J Clin Invest. 2008;118:1899–1910. doi: 10.1172/JCI34192. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Bevan S, Andersson DA. TRP channel antagonists for pain-opportunities beyond TRPV1. Curr Opin Investig Drugs. 2009;10:655–663. [PubMed] [Google Scholar]
  13. Binshtok AM, Wang H, Zimmermann K, Amaya F, Vardeh D, Shi L, et al. Nociceptors are interleukin-1beta sensors. J Neurosci. 2008;28:14062–14073. doi: 10.1523/JNEUROSCI.3795-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Birder LA, Kanai AJ, de Groat WC, Kiss S, Nealen ML, Burke NE, et al. Vanilloid receptor expression suggests a sensory role for urinary bladder epithelial cells. Proc Natl Acad Sci U S A. 2001;98:13396–13401. doi: 10.1073/pnas.231243698. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Brain SD, Williams TJ, Tippins JR, Morris HR, MacIntyre I. Calcitonin gene-related peptide is a potent vasodilator. Nature. 1985;313:54–56. doi: 10.1038/313054a0. [DOI] [PubMed] [Google Scholar]
  16. Brand L, Berman E, Schwen R, Loomans M, Janusz J, Bohne R, et al. NE-19550: a novel, orally active anti-inflammatory analgesic. Drugs Exp Clin Res. 1987;13:259–265. [PubMed] [Google Scholar]
  17. Caterina MJ, Schumacher MA, Tominaga M, Rosen TA, Levine JD, Julius D. The capsaicin receptor: a heat-activated ion channel in the pain pathway. Nature. 1997;389:816–824. doi: 10.1038/39807. [DOI] [PubMed] [Google Scholar]
  18. Caterina MJ, Leffler A, Malmberg AB, Martin WJ, Trafton J, Petersen-Zeitz KR, et al. Impaired nociception and pain sensation in mice lacking the capsaicin receptor. Science. 2000;288:306–313. doi: 10.1126/science.288.5464.306. [DOI] [PubMed] [Google Scholar]
  19. Cavanaugh DJ, Chesler AT, Jackson AC, Sigal YM, Yamanaka H, Grant R, et al. Trpv1 reporter mice reveal highly restricted brain distribution and functional expression in arteriolar smooth muscle cells. J Neurosci. 2011;31:5067–5077. doi: 10.1523/JNEUROSCI.6451-10.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Chao LK, Hua KF, Hsu HY, Cheng SS, Lin IF, Chen CJ, et al. Cinnamaldehyde inhibits pro-inflammatory cytokines secretion from monocytes/macrophages through suppression of intracellular signaling. Food Chem Toxicol. 2008;46:220–231. doi: 10.1016/j.fct.2007.07.016. [DOI] [PubMed] [Google Scholar]
  21. Chu CJ, Huang SM, De Petrocellis L, Bisogno T, Ewing SA, Miller JD, et al. N-oleoyldopamine, a novel endogenous capsaicin-like lipid that produces hyperalgesia. J Biol Chem. 2003;278:13633–13639. doi: 10.1074/jbc.M211231200. [DOI] [PubMed] [Google Scholar]
  22. Chuang HH, Prescott ED, Kong H, Shields S, Jordt SE, Basbaum AI, et al. Bradykinin and nerve growth factor release the capsaicin receptor from PtdIns(4,5)P2-mediated inhibition. Nature. 2001;411:957–962. doi: 10.1038/35082088. [DOI] [PubMed] [Google Scholar]
  23. Clapham DE. SnapShot: mammalian TRP channels. Cell. 2007;129:220. doi: 10.1016/j.cell.2007.03.034. [DOI] [PubMed] [Google Scholar]
  24. Corey DP, García-Añoveros J, Holt JR, Kwan KY, Lin SY, Vollrath MA, et al. TRPA1 is a candidate for the mechanosensitive transduction channel of vertebrate hair cells. Nature. 2004;432:723–730. doi: 10.1038/nature03066. [DOI] [PubMed] [Google Scholar]
  25. Cortright DN, Szallasi A. TRP channels and pain. Curr Pharm Des. 2009;15:1736–1749. doi: 10.2174/138161209788186308. [DOI] [PubMed] [Google Scholar]
  26. Dai Y, Wang S, Tominaga M, Yamamoto S, Fukuoka T, Higashi T. Sensitization of TRPA1 by PAR2 contributes to the sensation of inflammatory pain. J Clin Invest. 2007;117:1979–1987. doi: 10.1172/JCI30951. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Dib B. Effects of intracerebroventricular capsaicin on thermoregulatory behavior in the rat. Pharmacol Biochem Behav. 1982;16:23–27. doi: 10.1016/0091-3057(82)90007-7. [DOI] [PubMed] [Google Scholar]
  28. Doihara H, Nozawa K, Kawabata-Shoda E, Kojima R, Yokoyama T, Ito H. Molecular cloning and characterization of dog TRPA1 and AITC stimulate the gastrointestinal motility through TRPA1 in conscious dogs. Eur J Pharmacol. 2009;617:124–129. doi: 10.1016/j.ejphar.2009.06.038. [DOI] [PubMed] [Google Scholar]
  29. Earley S, Gonzales AL, Crnich R. Endothelium-dependent cerebral artery dilation mediated by TRPA1 and Ca2+-activated K+ channels. Circ Res. 2009;104:987–994. doi: 10.1161/CIRCRESAHA.108.189530. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Fernandes ES, Russell FA, Spina D, McDougall JJ, Graepel R, Gentry C, et al. A distinct role for transient receptor potential ankyrin 1, in addition to transient receptor potential vanilloid 1, in tumor necrosis factor α-induced inflammatory hyperalgesia and Freund's complete adjuvant-induced monarthritis. Arthritis Rheum. 2011;63:819–829. doi: 10.1002/art.30150. [DOI] [PubMed] [Google Scholar]
  31. Gavva NR, Bannon AW, Hovland DN, Jr, Lehto SG, Klionsky L, Surapaneni S, et al. Repeated administration of vanilloid receptor TRPV1 antagonists attenuates hyperthermia elicited by TRPV1 blockade. J Pharmacol Exp Ther. 2007a;323:128–137. doi: 10.1124/jpet.107.125674. [DOI] [PubMed] [Google Scholar]
  32. Gavva NR, Bannon AW, Surapaneni S, Hovland DN, Jr, Lehto SG, Gore A, et al. The vanilloid receptor TRPV1 is tonically activated in vivo and involved in body temperature regulation. J Neurosci. 2007b;27:3366–3374. doi: 10.1523/JNEUROSCI.4833-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Geppetti P, Nassini R, Materazzi S, Benemei S. The concept of neurogenic inflammation. BJU Int. 2008;101(Suppl. 3):2–6. doi: 10.1111/j.1464-410X.2008.07493.x. [DOI] [PubMed] [Google Scholar]
  34. Golech SA, McCarron RM, Chen Y, Bembry J, Lenz F, Mechoulam R, et al. Human brain endothelium: coexpression and function of vanilloid and endocannabinoid receptors. Brain Res Mol Brain Res. 2004;132:87–92. doi: 10.1016/j.molbrainres.2004.08.025. [DOI] [PubMed] [Google Scholar]
  35. Gröne A, Fonfara S, Baumgärtner W. Cell type-dependent cytokine expression after canine distemper virus infection. Viral Immunol. 2002;15:493–505. doi: 10.1089/088282402760312368. [DOI] [PubMed] [Google Scholar]
  36. Gu Q, Lin RL. Heavy metals zinc, cadmium, and copper stimulate pulmonary sensory neurons via direct activation of TRPA1. J Appl Physiol. 2010;108:891–897. doi: 10.1152/japplphysiol.01371.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Güler AD, Lee H, Iida T, Shimizu I, Tominaga M, Caterina M. Heat-evoked activation of the ion channel, TRPV4. J Neurosci. 2002;22:6408–6414. doi: 10.1523/JNEUROSCI.22-15-06408.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Hamamoto T, Takumida M, Hirakawa K, Takeno S, Tatsukawa T. Localization of transient receptor potential channel vanilloid subfamilies in the mouse larynx. Acta Otolaryngol. 2008;128:685–693. doi: 10.1080/00016480701669489. [DOI] [PubMed] [Google Scholar]
  39. Hamamoto T, Takumida M, Hirakawa K, Tatsukawa T, Ishibashi T. Localization of transient receptor potential vanilloid (TRPV) in the human larynx. Acta Otolaryngol. 2009;129:560–568. doi: 10.1080/00016480802273108. [DOI] [PubMed] [Google Scholar]
  40. Hayes P, Meadows HJ, Gunthorpe MJ, Harries MH, Duckworth DM, Cairns W, et al. Cloning and functional expression of a human orthologue of rat vanilloid receptor-1. Pain. 2000;88:205–215. doi: 10.1016/S0304-3959(00)00353-5. [DOI] [PubMed] [Google Scholar]
  41. Hsu CL, Yen GC. Effects of capsaicin on induction of apoptosis and inhibition of adipogenesis in 3T3-L1 cells. J Agric Food Chem. 2007;55:1730–1736. doi: 10.1021/jf062912b. [DOI] [PubMed] [Google Scholar]
  42. Hu H, Bandell M, Petrus MJ, Zhu MX, Patapoutian A. Zinc activates damage-sensing TRPA1 ion channels. Nat Chem Biol. 2009;5:183–190. doi: 10.1038/nchembio.146. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Hu H, Tian J, Zhu Y, Wang C, Xiao R, Herz JM, et al. Activation of TRPA1 channels by fenamate nonsteroidal anti-inflammatory drugs. Pflugers Arch. 2010;459:579–592. doi: 10.1007/s00424-009-0749-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Huang SM, Bisogno T, Trevisani M, Al-Hayani A, De Petrocellis L, Fezza F, et al. An endogenous capsaicin-like substance with high potency at recombinant and native vanilloid VR1 receptors. Proc Natl Acad Sci U S A. 2002;99:8400–8405. doi: 10.1073/pnas.122196999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Hwang JT, Park IJ, Shin JI, Lee YK, Lee SK, Baik HW, et al. Genistein, EGCG, and capsaicin inhibit adipocyte differentiation process via activating AMP-activated protein kinase. Biochem Biophys Res Commun. 2005;338:694–699. doi: 10.1016/j.bbrc.2005.09.195. [DOI] [PubMed] [Google Scholar]
  46. Hwang SW, Cho H, Kwak J, Lee SY, Kang CJ, Jung J, et al. Direct activation of capsaicin receptors by products of lipoxygenases: endogenous capsaicin-like substances. Proc Natl Acad Sci U S A. 2000;97:6155–6160. doi: 10.1073/pnas.97.11.6155. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Inoue K, Koizumi S, Fuziwara S, Denda S, Inoue K, Denda M. Functional vanilloid receptors in cultured normal human epidermal keratinocytes. Biochem Biophys Res Commun. 2002;291:124–129. doi: 10.1006/bbrc.2002.6393. [DOI] [PubMed] [Google Scholar]
  48. Jancsô G, Wollemann M. The effect of capsaicin on the adenylate cyclase activity of rat brain. Brain Res. 1977;123:323–329. doi: 10.1016/0006-8993(77)90483-8. [DOI] [PubMed] [Google Scholar]
  49. Jancsó N, Jancsó-Gábor A, Szolcsányi J. Direct evidence for neurogenic inflammation and its prevention by denervation and by pretreatment with capsaicin. Br J Pharmacol Chemother. 1967;31:138–151. doi: 10.1111/j.1476-5381.1967.tb01984.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Jancsó-Gábor A, Szolcsányi J, Jancsó N. Irreversible impairment of thermoregulation induced by capsaicin and similar pungent substances in rats and guinea-pigs. J Physiol. 1970a;206:495–507. doi: 10.1113/jphysiol.1970.sp009027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Jancsó-Gábor A, Szolcsányi J, Jancsó N. Stimulation and desensitization of the hypothalamic heat-sensitive structures by capsaicin in rats. J Physiol. 1970b;208:449–459. doi: 10.1113/jphysiol.1970.sp009130. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Jaquemar D, Schenker T, Trueb B. An ankyrin-like protein with transmembrane domains is specifically lost after oncogenic transformation of human fibroblasts. J Biol Chem. 1999;274:7325–7333. doi: 10.1074/jbc.274.11.7325. [DOI] [PubMed] [Google Scholar]
  53. Jin K, Xie L, Kim SH, Parmentier-Batteur S, Sun Y, Mao XO, et al. Defective adult neurogenesis in CB1 cannabinoid receptor knockout mice. Mol Pharmacol. 2004;66:204–208. doi: 10.1124/mol.66.2.204. [DOI] [PubMed] [Google Scholar]
  54. Jordt SE, Bautista DM, Chuang HH, McKemy DD, Zygmunt PM, Högestätt ED, et al. Mustard oils and cannabinoids excite sensory nerve fibres through the TRP channel ANKTM1. Nature. 2004;427:260–265. doi: 10.1038/nature02282. [DOI] [PubMed] [Google Scholar]
  55. Kang JH, Goto T, Han IS, Kawada T, Kim YM, Yu R. Dietary capsaicin reduces obesity-induced insulin resistance and hepatic steatosis in obese mice fed a high-fat diet. Obesity (Silver Spring) 2010;18:780–787. doi: 10.1038/oby.2009.301. [DOI] [PubMed] [Google Scholar]
  56. Karashima Y, Talavera K, Everaerts W, Janssens A, Kwan KY, Vennekens R, et al. TRPA1 acts as a cold sensor in vitro and in vivo. Proc Natl Acad Sci U S A. 2009;106:1273–1278. doi: 10.1073/pnas.0808487106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Kark T, Bagi Z, Lizanecz E, Pásztor ET, Erdei N, Czikora A, et al. Tissue-specific regulation of microvascular diameter: opposite functional roles of neuronal and smooth muscle located vanilloid receptor-1. Mol Pharmacol. 2008;73:1405–1412. doi: 10.1124/mol.107.043323. [DOI] [PubMed] [Google Scholar]
  58. Kawabata F, Inoue N, Masamoto Y, Matsumura S, Kimura W, Kadowaki M, et al. Non-pungent capsaicin analogs (capsinoids) increase metabolic rate and enhance thermogenesis via gastrointestinal TRPV1 in mice. Biosci Biotechnol Biochem. 2009;73:2690–2697. doi: 10.1271/bbb.90555. [DOI] [PubMed] [Google Scholar]
  59. Kaznacheyeva E, Glushankova L, Bugaj V, Zimina O, Skopin A, Alexeenko V, et al. Suppression of TRPC3 leads to disappearance of store-operated channels and formation of a new type of store-independent channels in A431 cells. J Biol Chem. 2007;282:23655–23662. doi: 10.1074/jbc.M608378200. [DOI] [PubMed] [Google Scholar]
  60. Keeble JE, Brain SD. Capsaicin-induced vasoconstriction in the mouse knee joint: a study using TRPV1 knockout mice. Neurosci Lett. 2006;401:55–58. doi: 10.1016/j.neulet.2006.02.083. [DOI] [PubMed] [Google Scholar]
  61. Kim SJ, Lee SA, Yun SJ, Kim JK, Park JS, Jeong HS, et al. Expression of vanilloid receptor 1 in cultured fibroblast. Exp Dermatol. 2006;15:362–367. doi: 10.1111/j.0906-6705.2006.00418.x. [DOI] [PubMed] [Google Scholar]
  62. Kochukov MY, McNearney TA, Fu Y, Westlund KN. Thermosensitive TRP ion channels mediate cytosolic calcium response in human synoviocytes. Am J Physiol Cell Physiol. 2006;291:C424–C432. doi: 10.1152/ajpcell.00553.2005. [DOI] [PubMed] [Google Scholar]
  63. Koh WS, Yoon SY, Kwon BM, Jeong TC, Nam KS, Han MY. Cinnamaldehyde inhibits lymphocyte proliferation and modulates T-cell differentiation. Int J Immunopharmacol. 1998;20:643–660. doi: 10.1016/s0192-0561(98)00064-2. [DOI] [PubMed] [Google Scholar]
  64. Koltzenburg M, McMahon SB. Plasma extravasation in the rat urinary bladder following mechanical, electrical and chemical stimuli: evidence for a new population of chemosensitive primary sensory afferents. Neurosci Lett. 1986;72:352–356. doi: 10.1016/0304-3940(86)90540-9. [DOI] [PubMed] [Google Scholar]
  65. Kwan KY, Allchorne AJ, Vollrath MA, Christensen AP, Zhang DS, Woolf CJ, et al. TRPA1 contributes to cold, mechanical, and chemical nociception but is not essential for hair-cell transduction. Neuron. 2006;50:277–289. doi: 10.1016/j.neuron.2006.03.042. [DOI] [PubMed] [Google Scholar]
  66. Kwan KY, Glazer JM, Corey DP, Rice FL, Stucky CL. TRPA1 modulates mechanotransduction in cutaneous sensory neurons. J Neurosci. 2009;29:4808–4819. doi: 10.1523/JNEUROSCI.5380-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Latorre R, Zaelzer C, Brauchi S. Structure-functional intimacies of transient receptor potential channels. Q Rev Biophys. 2009;42:201–246. doi: 10.1017/S0033583509990072. [DOI] [PubMed] [Google Scholar]
  68. Lee SH, Lee SY, Son DJ, Lee H, Yoo HS, Song S, et al. Inhibitory effect of 2′-hydroxycinnamaldehyde on nitric oxide production through inhibition of NF-κB activation in RAW 264.7 cells. Biochem Pharmacol. 2005;69:791–799. doi: 10.1016/j.bcp.2004.11.013. [DOI] [PubMed] [Google Scholar]
  69. Li J, Wang DH. Function and regulation of the vanilloid receptor in rats fed a high salt diet. J Hypertens. 2003;21:1–6. doi: 10.1097/00004872-200308000-00017. [DOI] [PubMed] [Google Scholar]
  70. Liedtke W, Tobin DM, Bargmann CI, Friedman JM. Mammalian TRPV4 (VR-OAC) directs behavioral responses to osmotic and mechanical stimuli in Caenorhabditis elegans. Proc Natl Acad Sci U S A. 2003;100(Suppl. 2):14531–14536. doi: 10.1073/pnas.2235619100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Luo D, Zhang YW, Peng WJ, Peng J, Chen QQ, Li D, et al. Transient receptor potential vanilloid 1-mediated expression and secretion of endothelial cell-derived calcitonin gene-related peptide. Regul Pept. 2008;150:66–72. doi: 10.1016/j.regpep.2008.05.007. [DOI] [PubMed] [Google Scholar]
  72. Macpherson LJ, Geierstanger BH, Viswanath V, Bandell M, Eid SR, Hwang S, et al. The pungency of garlic: activation of TRPA1 and TRPV1 in response to allicin. Curr Biol. 2005;15:929–934. doi: 10.1016/j.cub.2005.04.018. [DOI] [PubMed] [Google Scholar]
  73. Macpherson LJ, Hwang SW, Miyamoto T, Dubin AE, Patapoutian A, Story GM. More than cool: promiscuous relationships of menthol and other sensory compounds. Mol Cell Neurosci. 2006;32:335–343. doi: 10.1016/j.mcn.2006.05.005. [DOI] [PubMed] [Google Scholar]
  74. Maroto R, Raso A, Wood TG, Kurosky A, Martinac B, Hamill OP. TRPC1 forms the stretch-activated cation channel in vertebrate cells. Nat Cell Biol. 2005;7:179–185. doi: 10.1038/ncb1218. [DOI] [PubMed] [Google Scholar]
  75. Marsh SJ, Stansfeld CE, Brown DA, Davey R, McCarthy D. The mechanism of action of capsaicin on sensory C-type neurons and their axons in vitro. Neuroscience. 1987;23:275–289. doi: 10.1016/0306-4522(87)90289-2. [DOI] [PubMed] [Google Scholar]
  76. Masamoto Y, Kawabata F, Fushiki T. Intragastric administration of TRPV1, TRPV3, TRPM8, and TRPA1 agonists modulates autonomic thermoregulation in different manners in mice. Biosci Biotechnol Biochem. 2009;73:1021–1027. doi: 10.1271/bbb.80796. [DOI] [PubMed] [Google Scholar]
  77. Masuda Y, Haramizu S, Oki K, Ohnuki K, Watanabe T, Yazawa S, et al. Upregulation of uncoupling proteins by oral administration of capsiate, a nonpungent capsaicin analog. J Appl Physiol. 2003;95:2408–2415. doi: 10.1152/japplphysiol.00828.2002. [DOI] [PubMed] [Google Scholar]
  78. McNamara CR, Mandel-Brehm J, Bautista DM, Siemens J, Deranian KL, Zhao M, et al. TRPA1 mediates formalin-induced pain. Proc Natl Acad Sci U S A. 2007;104:13525–13530. doi: 10.1073/pnas.0705924104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Mergler S, Valtink M, Coulson-Thomas VJ, Lindemann D, Reinach PS, Engelmann K, et al. TRPV channels mediate temperature-sensing in human corneal endothelial cells. Exp Eye Res. 2010;90:758–770. doi: 10.1016/j.exer.2010.03.010. [DOI] [PubMed] [Google Scholar]
  80. Mezey E, Tóth ZE, Cortright DN, Arzubi MK, Krause JE, Elde R, et al. Distribution of mRNA for vanilloid receptor subtype 1 (VR1), and VR1-like immunoreactivity, in the central nervous system of the rat and human. Proc Natl Acad Sci U S A. 2000;97:3655–3660. doi: 10.1073/pnas.060496197. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Nagata K, Duggan A, Kumar G, García-Añoveros J. Nociceptor and hair cell transducer properties of TRPA1, a channel for pain and hearing. J Neurosci. 2005;25:4052–4061. doi: 10.1523/JNEUROSCI.0013-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Nozawa K, Kawabata-Shoda E, Doihara H, Kojima R, Okada H, Mochizuki S, et al. TRPA1 regulates gastrointestinal motility through serotonin release from enterochromaffin cells. Proc Natl Acad Sci U S A. 2009;106:3408–3413. doi: 10.1073/pnas.0805323106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Ohnuki K, Haramizu S, Oki K, Watanabe T, Yazawa S, Fushiki T. Administration of capsiate, a non-pungent capsaicin analog, promotes energy metabolism and suppresses body fat accumulation in mice. Biosci Biotechnol Biochem. 2001;65:2735–2740. doi: 10.1271/bbb.65.2735. [DOI] [PubMed] [Google Scholar]
  84. Pedersen SF, Owsianik G, Nilius B. TRP channels: an overview. Cell Calcium. 2005;38:233–252. doi: 10.1016/j.ceca.2005.06.028. [DOI] [PubMed] [Google Scholar]
  85. Peier AM, Moqrich A, Hergarden AC, Reeve AJ, Andersson DA, Story GM, et al. A TRP channel that senses cold stimuli and menthol. Cell. 2002;108:705–715. doi: 10.1016/s0092-8674(02)00652-9. [DOI] [PubMed] [Google Scholar]
  86. Pozsgai G, Bodkin JV, Graepel R, Bevan S, Andersson DA, Brain SD. Evidence for the pathophysiological relevance of TRPA1 receptors in the cardiovascular system in vivo. Cardiovasc Res. 2010;87:760–768. doi: 10.1093/cvr/cvq118. [DOI] [PubMed] [Google Scholar]
  87. Riccio A, Mattei C, Kelsell RE, Medhurst AD, Calver AR, Randall AD, et al. Cloning and functional expression of human short TRP7, a candidate protein for store-operated Ca2+ influx. J Biol Chem. 2002;277:12302–12309. doi: 10.1074/jbc.M112313200. [DOI] [PubMed] [Google Scholar]
  88. Romanovsky AA, Almeida MC, Garami A, Steiner AA, Norman MH, Morrison SF, et al. The transient receptor potential vanilloid-1 channel in thermoregulation: a thermosensor it is not. Pharmacol Rev. 2009;61:228–261. doi: 10.1124/pr.109.001263. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Ruparel NB, Patwardhan AM, Akopian AN, Hargreaves KM. Homologous and heterologous desensitization of capsaicin and mustard oil responses utilize different cellular pathways in nociceptors. Pain. 2008;135:271–279. doi: 10.1016/j.pain.2007.06.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Salas MM, Hargreaves KM, Akopian AN. TRPA1-mediated responses in trigeminal sensory neurons: interaction between TRPA1 and TRPV1. Eur J Neurosci. 2009;29:1568–1578. doi: 10.1111/j.1460-9568.2009.06702.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Sasamura T, Sasaki M, Tohda C, Kuraishi Y. Existence of capsaicin-sensitive glutamatergic terminals in rat hypothalamus. Neuroreport. 1998;9:2045–2048. doi: 10.1097/00001756-199806220-00025. [DOI] [PubMed] [Google Scholar]
  92. Saunders CI, Kunde DA, Crawford A, Geraghty DP. Expression of transient receptor potential vanilloid 1 (TRPV1) and 2 (TRPV2) in human peripheral blood. Mol Immunol. 2007;44:1429–1435. doi: 10.1016/j.molimm.2006.04.027. [DOI] [PubMed] [Google Scholar]
  93. Savidge J, Davis C, Shah K, Colley S, Phillips E, Ranasinghe S, et al. Cloning and functional characterization of the guinea pig vanilloid receptor 1. Neuropharmacology. 2002;43:450–456. doi: 10.1016/s0028-3908(02)00122-3. [DOI] [PubMed] [Google Scholar]
  94. Schaible HG, Schmidt RF. Excitation and sensitization of fine articular afferents from cat's knee joint by prostaglandin E2. J Physiol. 1988;403:91–104. doi: 10.1113/jphysiol.1988.sp017240. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Seki N, Shirasaki H, Kikuchi M, Sakamoto T, Watanabe N, Himi T. Expression and localization of TRPV1 in human nasal mucosa. Rhinology. 2006;44:128–134. [PubMed] [Google Scholar]
  96. Smart D, Gunthorpe MJ, Jerman JC, Nasir S, Gray J, Muir AI, et al. The endogenous lipid anandamide is a full agonist at the human vanilloid receptor (hVR1) Br J Pharmacol. 2000;129:227–230. doi: 10.1038/sj.bjp.0703050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Southall MD, Li T, Gharibova LS, Pei Y, Nicol GD, Travers JB. Activation of epidermal vanilloid receptor-1 induces release of proinflammatory mediators in human keratinocytes. J Pharmacol Exp Ther. 2003;304:217–222. doi: 10.1124/jpet.102.040675. [DOI] [PubMed] [Google Scholar]
  98. Staruschenko A, Jeske NA, Akopian AN. Contribution of TRPV1-TRPA1 interaction to the single channel properties of the TRPA1 channel. J Biol Chem. 2010;285:15167–15177. doi: 10.1074/jbc.M110.106153. [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Steiner AA, Turek VF, Almeida MC, Burmeister JJ, Oliveira DL, Roberts JL, et al. Nonthermal activation of transient receptor potential vanilloid-1 channels in abdominal viscera tonically inhibits autonomic cold-defense effectors. J Neurosci. 2007;27:7459–7468. doi: 10.1523/JNEUROSCI.1483-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Story GM, Peier AM, Reeve AJ, Eid SR, Mosbacher J, Hricik TR, et al. ANKTM1, a TRP-like channel expressed in nociceptive neurons, is activated by cold temperatures. Cell. 2003;112:819–829. doi: 10.1016/s0092-8674(03)00158-2. [DOI] [PubMed] [Google Scholar]
  101. Streng T, Axelsson HE, Hedlund P, Andersson DA, Jordt SE, Bevan S, et al. Distribution and function of the hydrogen sulfide-sensitive TRPA1 ion channel in rat urinary bladder. Eur Urol. 2008;53:391–399. doi: 10.1016/j.eururo.2007.10.024. [DOI] [PubMed] [Google Scholar]
  102. Stucky CL, Dubin AE, Jeske NA, Malin SA, McKemy DD, Story GM. Roles of transient receptor potential channels in pain. Brain Res Rev. 2009;60:2–23. doi: 10.1016/j.brainresrev.2008.12.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Szallasi A, Blumberg PM. Resiniferatoxin, a phorbol-related diterpene, acts as an ultrapotent analog of capsaicin, the irritant constituent in red pepper. Neuroscience. 1989;30:515–520. doi: 10.1016/0306-4522(89)90269-8. [DOI] [PubMed] [Google Scholar]
  104. Szallasi A, Nilsson S, Farkas-Szallasi T, Blumberg PM, Hokfelt T, Lundberg JM. Vanilloid (capsaicin) receptors in the rat: distribution in the brain, regional differences in the spinal cord, axonal transport to the periphery, and depletion by systemic vanilloid treatment. Brain Res. 1995;703:175–183. doi: 10.1016/0006-8993(95)01094-7. [DOI] [PubMed] [Google Scholar]
  105. Szolcsányi J, Jancsó-Gábor A. Sensory effects of capsaicin congeners I. Relationship between chemical structure and pain-producing potency of pungent agents. Arzneimittelforschung. 1975;25:1877–1881. [PubMed] [Google Scholar]
  106. Szolcsányi J, Jancśo-Gábor A, Joo F. Functional and fine structural characteristics of the sensory neuron blocking effect of capsaicin. Naunyn Schmiedebergs Arch Pharmacol. 1975;287:157–169. doi: 10.1007/BF00510447. [DOI] [PubMed] [Google Scholar]
  107. Szolcsányi J, Szallasi A, Szallasi Z, Joo F, Blumberg PM. Resiniferatoxin: an ultrapotent selective modulator of capsaicin-sensitive primary afferent neurons. J Pharmacol Exp Ther. 1990;255:923–928. [PubMed] [Google Scholar]
  108. Tamayo N, Liao H, Stec MM, Wang X, Chakrabarti P, Retz D, et al. Design and synthesis of peripherally restricted transient receptor potential vanilloid 1 (TRPV1) antagonists. J Med Chem. 2008;51:2744–2757. doi: 10.1021/jm7014638. [DOI] [PubMed] [Google Scholar]
  109. Taylor-Clark TE, Undem BJ, Macglashan DW, Jr, Ghatta S, Carr MJ, McAlexander MA. Prostaglandin-induced activation of nociceptive neurons via direct interaction with transient receptor potential A1 (TRPA1) Mol Pharmacol. 2008;73:274–281. doi: 10.1124/mol.107.040832. [DOI] [PubMed] [Google Scholar]
  110. Tominaga M, Caterina MJ, Malmberg AB, Rosen TA, Gilbert H, Skinner K. The cloned capsaicin receptor integrates multiple pain-producing stimuli. Neuron. 1998;21:531–543. doi: 10.1016/s0896-6273(00)80564-4. [DOI] [PubMed] [Google Scholar]
  111. Tóth DM, Szoke E, Bölcskei K, Kvell K, Bender B, Bosze Z, et al. Nociception, neurogenic inflammation and thermoregulation in TRPV1 knockdown transgenic mice. Cell Mol Life Sci. 2011;68:2589–2601. doi: 10.1007/s00018-010-0569-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Trevisani M, Siemens J, Materazzi S, Bautista DM, Nassini R, Campi B, et al. 4-Hydroxynonenal, an endogenous aldehyde, causes pain and neurogenic inflammation through activation of the irritant receptor TRPA1. Proc Natl Acad Sci U S A. 2007;104:13519–13524. doi: 10.1073/pnas.0705923104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Tsutsumi M, Denda S, Ikeyama K, Goto M, Denda M. Exposure to low temperature induces elevation of intracellular calcium in cultured human keratinocytes. J Invest Dermatol. 2010;130:1945–1948. doi: 10.1038/jid.2010.33. [DOI] [PubMed] [Google Scholar]
  114. Walker KM, Urban L, Medhurst SJ, Patel S, Panesar M, Fox AJ, et al. The VR1 antagonist capsazepine reverses mechanical hyperalgesia in models of inflammatory and neuropathic pain. J Pharmacol Exp Ther. 2003;304:56–62. doi: 10.1124/jpet.102.042010. [DOI] [PubMed] [Google Scholar]
  115. Wang L, Wang DH. TRPV1 gene knockout impairs postischemic recovery in isolated perfused heart in mice. Circulation. 2005;112:3617–3623. doi: 10.1161/CIRCULATIONAHA.105.556274. [DOI] [PubMed] [Google Scholar]
  116. Wang S, Dai Y, Fukuoka T, Yamanaka H, Kobayashi K, Obata K, et al. Phospholipase C and protein kinase A mediate bradykinin sensitization of TRPA1: a molecular mechanism of inflammatory pain. Brain. 2008;131:1241–1251. doi: 10.1093/brain/awn060. [DOI] [PubMed] [Google Scholar]
  117. Wang YX, Wang J, Wang C, Liu J, Shi LP, Xu M, et al. Functional expression of transient receptor potential vanilloid-related channels in chronically hypoxic human pulmonary arterial smooth muscle cells. J Membr Biol. 2008;223:151–159. doi: 10.1007/s00232-008-9121-9. [DOI] [PubMed] [Google Scholar]
  118. Xu H, Blair NT, Clapham DE. Camphor activates and strongly desensitizes the transient receptor potential vanilloid subtype 1 channel in a vanilloid-independent mechanism. J Neurosci. 2005;25:8924–8937. doi: 10.1523/JNEUROSCI.2574-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  119. Yang D, Luo Z, Ma S, Wong WT, Ma L, Zhong J, et al. Activation of TRPV1 by dietary capsaicin improves endothelium-dependent vasorelaxation and prevents hypertension. Cell Metab. 2010;12:130–141. doi: 10.1016/j.cmet.2010.05.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Yang XR, Lin MJ, McIntosh LS, Sham JS. Functional expression of transient receptor potential melastatin- and vanilloid-related channels in pulmonary arterial and aortic smooth muscle. Am J Physiol Lung Cell Mol Physiol. 2006;290:L1267–L1276. doi: 10.1152/ajplung.00515.2005. [DOI] [PubMed] [Google Scholar]
  121. Yao X, Garland CJ. Recent developments in vascular endothelial cell transient receptor potential channels. Circ Res. 2005;97:853–863. doi: 10.1161/01.RES.0000187473.85419.3e. [DOI] [PubMed] [Google Scholar]
  122. Zhang F, Yang H, Wang Z, Mergler S, Liu H, Kawakita T, et al. Transient receptor potential vanilloid 1 activation induces inflammatory cytokine release in corneal epithelium through MAPK signaling. J Cell Physiol. 2007a;213:730–739. doi: 10.1002/jcp.21141. [DOI] [PubMed] [Google Scholar]
  123. Zhang LL, Yan Liu D, Ma LQ, Luo ZD, Cao TB, Zhong J, et al. Activation of transient receptor potential vanilloid type-1 channel prevents adipogenesis and obesity. Circ Res. 2007b;100:1063–1070. doi: 10.1161/01.RES.0000262653.84850.8b. [DOI] [PubMed] [Google Scholar]
  124. Zygmunt PM, Petersson J, Andersson DA, Chuang H, Sørgård M, Di Marzo V, et al. Vanilloid receptors on sensory nerves mediate the vasodilator action of anandamide. Nature. 1999;400:452–457. doi: 10.1038/22761. [DOI] [PubMed] [Google Scholar]

Articles from British Journal of Pharmacology are provided here courtesy of The British Pharmacological Society

RESOURCES