3,789
Views
46
CrossRef citations to date
0
Altmetric
Research Article

Sensing hot and cold with TRP channels

Pages 388-398 | Received 30 Jul 2010, Accepted 11 Jan 2011, Published online: 18 May 2011

Abstract

The past decade has witnessed the cloning of a new family of ion channels that are responsive to temperature. Six of these transient receptor potential (TRP) channels are proposed to be involved in thermosensation and are located in sensory nerves and skin. The TRPV1, TRPV2, TRPV3, and TRPV4 channels have incompletely overlapping functions over a broad thermal range from warm to hot. Deletion of the individual TRPV1, TRPV3, and TRPV4 channels in mice has established their physiological role in thermosensation. In all cases thermosensation is not completely abolished – suggesting some functional redundancy among the channels. Notably, the TRPV2 channel is responsive to hot temperatures in heterologous systems, but its physiological relevance in vivo has not been established. Cool and cold temperatures are sensed by TRPM8 and TRPA1 family members. Currently, the pharmaceutical industry is developing agonists and antagonists for the various TRP channels. For instance, TRPV1 receptor agonists produce hypothermia, while antagonists induce hyperthermia. Recent investigations have found that different regions of the TRPV1 receptor are responsive to temperature, nociceptive stimuli, and various chemical agents. With this information, it has been possible to develop a TRPV1 compound that blocks responses to capsaicin and acid while leaving temperature sensitivity intact. These channels have important implications for hyperthermia research and may help to identify previously unexplored mechanisms in different tissues that are responsive to thermal stress.

Introduction

Maintenance of body temperature is a basic physiological process for many different organisms. Mammals defend their body temperatures through a variety of physiological and behavioral responses which help them to maintain temperature within a narrow range that is optimal for ongoing biochemical processes. Innate acute nocifensive behaviors permit escape from environmental conditions of intense cold or heat so that tissue damage, morbidity, or mortality do not ensue. Responses to temperature change require the organism to possess sensors that can detect thermal changes in the environment. Very recent work has identified a number of receptors in the skin that respond to temperature changes in the environment. Six different thermo-sensitive receptors have been newly identified to play important roles in thermoregulation and these receptors are members of the transient receptor potential (TRP) ion channel family and have been termed thermo-TRP channels ().

Table I.  Characteristics of thermo-TRP channels.

TRP channels that sense heat

TRP channels were first identified in a phototransduction mutant in Drosophila that showed a transient instead of a sustained response to light Citation[1]. Since that time, the first mammalian TRP channel was found to reside in primary sensory neurons and it was termed TRPV1 (TRP vanilloid receptor type 1) Citation[2–4] (). Four identical TRPV1 subunits, each containing six transmembrane domains with a pore loop between domains 5 and 6, assemble to form the channel Citation[5]. In heterologous systems this channel is weakly Ca2+-selective, and when stimulated by capsaicin, resiniferatoxin, acid, endocannabinoids, certain eicosanoids, and heat (>43°C), it produces outwardly rectifying cation currents Citation[2–4], Citation[6]. As an aside, in avian species TRPV1 responds to heat but not capsaicin Citation[7]. Since in whole-cell recordings, excised patches with TRPV1 can be activated by increased temperature, it is likely that this receptor is stimulated directly by the proton mode of the receptor Citation[6], Citation[8]. TRPV1 is expressed primarily in small- to medium-diameter peptidergic and non-peptidergic sensory neurons within the dorsal root, trigeminal, and nodose sensory ganglia that are characteristic of nociceptive C- and Aδ-fibers Citation[2], Citation[6]. It is also found within brain Citation[9], Citation[10]. In the periphery, termini of capsaicin-sensitive neurons release the neuropeptides, substance P and calcitonin gene-related peptide that can lead to neurogenic inflammation Citation[11]. Recently, two splice variants of TRPV1 have been identified Citation[12]. When murine TRPV1β is expressed with TRPV1α channels, it exerts dominant negative actions; the human variant is stimulated by acid and capsaicin Citation[13]. Additional TRPV1 mRNA species have been visualized in kidney and brain Citation[14], Citation[15]. TRPV1-knockout mice have been generated and electrophysiological responses to capsaicin, acid, and heat (43°C) were reduced or absent in these mutants, while responses to >50–55°C were conserved Citation[16], Citation[17]. In tail immersion assay, withdrawal latencies at temperatures >48°C were protracted for TRPV1-knockout animals, but they had normal latencies at lower temperatures Citation[16]. A similar relationship was seen with the hot-plate assay Citation[16]. Despite this result, another group was unable to detect any genotype differences in the hot-plate or radiant heat assays Citation[17]. Nonetheless, inflammation-induced heat hyperalgesia was virtually abolished in TRPV1-knockout animals Citation[16]. When capsaicin was injected into the plantar skin of the hind-paw, robust licking and paw shaking were observed in wild type mice Citation[16]. By comparison, little or no response was evident and edema was less apparent in the knockouts. In a skin-nerve preparation ∼50% of small dorsal root ganglion cells from wild type mice were responsive to heat stimuli up to 47°C, only 16% of cells from knockout mice were activated by heat Citation[16]. Nevertheless, no genotype differences were observed in thermal thresholds for units that were stimulated by heat stimuli; however, the mean evoked discharge was decreased by ∼45% in units from mutant mice. Interestingly, peripheral injection of capsaicin was sufficient to lower body temperature by ∼6°C in wild type animals and it took up to 2 hours for recovery; TRPV1-knockout mice were completely unresponsive. Additional experiments with TRPV1-knockout mice have revealed that they are deficient in mounting a fever response to lipopolysaccharide Citation[18]. Although these collective findings strongly support a role for TRPV1 receptors in thermal responses, they cannot account for all responses to heat.

Effects of TRPV1 on temperature regulation have been investigated in more detail. Capsaicin and resiniferatoxin are known to produce hypothermia Citation[19], Citation[20]. As described above, body temperatures of TRPV1-knockout mice were not affected by capsaicin administration Citation[16]. Subsequent studies with mutants have shown that the circadian rhythm for temperature, tolerance to decreased (4°C) or increased (35°C) environmental temperatures, and responses to ethanol-induced hyperthermia were normal Citation[18]. However, fever production by lipopolysaccharide was attenuated in TRPV1-knockout mice. In a separate investigation Citation[21], the amplitude of daily body temperatures were higher for TRPV1-knockout animals than for wild type controls. As a control, wild type mice were treated with capsaicin for 2 consecutive days to down-regulate the receptor Citation[21]. Interestingly, the amplitude of the daily body temperatures was even higher than that in the knockout mice. When the TRPV1 mice were placed into a 37°C environment for 30 min, body temperatures were increased to the same extent in wild type and mutant animals; the capsaicin-desensitized animals had to be removed from the conditions after 15 min because of development of imminent heat-stroke (i.e., their body temperatures reached 42°C). These results indicate that TRPV1-knockout animals can defend their body temperatures against acute overheating. Notably, these experiments also reveal that down-regulation of the channel more severely effects temperature regulation than deletion of the gene – possibly due to off-target effects in the intact animal and compensatory effects in the mutant mouse. Importantly, this difference in response has important implications for the development of compounds to be used for pharmacotherapy.

Additional temperature regulation experiments have been conducted with TRPV1 agonists and antagonists. TRPV1 receptor agonists cause hypothermia and shivering at least through skin vasodilation and reduced metabolism Citation[22]. Recent work with a variety of distinct TRPV1 receptor agonists and antagonists have found them to produce hypothermia and hyperthermia, respectively in rats, mice, dogs, monkeys, and humans, but not in TRPV1-knockout mice Citation[23]. The hyperthermia appears to be transient and tolerance to repeated drug administration develops, suggesting that other mechanisms are involved in maintaining homeostasis. Pharmacological antagonism of the TRPV1 channel also produces hyperthermia in humans Citation[24]. Together, these data indicate that TRPV1 is tonically active and that it plays a role in the regulation of body temperature through a variety of thermoeffector mechanisms in the periphery and central nervous systems [c.f., Citation[10], Citation[25].

TRPV2 has been found to mediate high-threshold heat sensations (>52°C) in vitro. Since it is expressed in medium- to large-diameter sensory neurons, it is thought to be associated with Aδ fibers Citation[3] (). TRPV2 is also found within the brain Citation[10], Citation[25]. In contrast to TRPV1 channels, TRPV2 is not responsive to capsaicin and its distribution across tissues is wider Citation[26]. Sensory C-fiber neurons from TRPV1-knockout mice that do not contain putative TRPV2 immunoreactivity respond normally to noxious heat Citation[27]. However, only one such neuron was identified in this study. Investigations by other researchers have localized TRPV2 primarily to myelinated fibers that were mechanically sensitive; no evidence for expression in C-fibers was observed Citation[28]. The few TRPV2-containing cells that did respond to heat also were active to cold stimuli. While TRPV-2 channels appear to be activated by heat in vitro, evidence that the channels mediate sensation to heat in vivo is lacking.

In heterologous systems TRPV3 is activated by camphor and increases in thermal temperatures (>33°C) Citation[29–32] (). As compared to other TRP channels, TRPV3 shows a biphasic activation profile Citation[33], Citation[34]. The initial sensitizing phase develops gradually and is followed abruptly by a secondary phase with a current of larger amplitude and the loss of the outer rectifying current. TRPV3 is expressed in keratinocytes, but not in sensory neurons Citation[30]. TRPV3 is also found within the brain Citation[20], Citation[25]. There is some evidence that TRPV3 may interact with TRPV1 and TRPV2 channels Citation[31], Citation[35]; however, the relationship between TRPV1 and TRPV3 channels is controversial Citation[35]. Disruption of TRPV3 exerts no effects on body weight, body temperature, exploratory activity in the open field, or anxiety-like responses Citation[29]. In a temperature preference test with a choice between two temperature zones, 92% of the wild type animals preferred to be on the 35°C plate, whereas only 64% of the mutants selected this plate over the room temperature one Citation[29]. Another thermotaxis assay run with both plates at room temperature or one at room temperature and the other at 15°C failed to distinguish responses between the genotypes Citation[29]. In a linear temperature gradient assay (15–55°C), wild type mice developed a preference for zones between 30–38°C within the first 25 minutes Citation[29]. Knockouts took 60 min to develop this preference. In a tail-immersion test, tail-flick was delayed in the TRPV3-knockout animals at 48, 50, and 52°C; a similar protracted response was seen with the hot-plate assay at 45, 50, and 55°C Citation[29]. Electrophysiological recordings with cultured keratinocytes showed that most wild type cells responded with gradually increasing current responses to 37°C pulses, whereas cells from knockout animals were unresponsive to this stimulus or displayed TRPV4-like desensitization responses Citation[29]. Hence, disruption of Trpv3 in mice serves to impair responses to noxious thermal stimuli, but it does not abolish them.

TRPV4 channels are also sensitive to heat stimuli at 25–34°C in heterologous systems Citation[36] (). TRPV4-knockout mice have been developed and these mutants show impairments in osmotic regulation or alterations in vasopressin secretion Citation[37], Citation[38]. TRPV4 is expressed in skin, endothelium, kidney, dorsal root ganglion, and brain Citation[37], Citation[38–40]. Importantly, TRPV4-knockout mice were deficient in responses to tail pressure and acid nociception; however, responses to radiant heating or to a hot plate (35–50°C) were not different from those of the wild type controls Citation[37], Citation[41], Citation[42]. Electrophysiological recordings revealed that ∼80% of the hind-paw femoral nerves of wild type mice were responsive to 40°C, while only ∼15% of nerves in TRPV4-knockout samples were activated Citation[42]. When temperature was increased from 25 to 50°C, nerves from wild type mice began responding at 33°C, whereas those from knockout animals did not respond until 40°C was reached Citation[42]. Notably, the numbers of heat-sensitive neurons and the magnitudes of evoked activities of the 40°C-sensitive neurons were similar for wild type and TRPV4-knockout mice. Additionally, no genotype differences were noted in responses at 50°C. Since inflammation or swelling can increase sensitivity to temperature Citation[40], animals were tested for hot-plate (40°C) responses before or 20 minutes after injection of carrageenan into the hind-paws Citation[42]. Despite there being no genotype differences in swelling to carrageenan, wild type latencies to escape to 40°C were decreased from baseline following injection while those for mutants were not distinguished between baseline and carrageenan conditions. Following carrageenan administration latencies to escape from 37.5 and 42.5°C were also prolonged for TRPV4-knockout mice, while both genotypes responded similarly to 50°C. Electrophysiology studies indicated that while carrageenan-induced discharges were different from the magnitude of those induced by temperature, the threshold for warmth-sensitive neural activity was reduced. Parenthetically, it should be emphasized that hypotonicity, such as swelling, can lead to production of 5′,6′-epoxyeicosatrienoic acid which can directly activate TRPV4 channels Citation[44], Citation[45]. These findings indicate that TRPV4 is responsive to thermal hyperalgesia at warm temperatures and that this channel is responsible for determining the sensitivity of the response. It should be emphasized, however, that heat may activate TRP4 through indirect mechanisms since in whole cell configurations in excised membrane patches TRPV4 is not activated by heat Citation[46].

In a separate set of experiments, Lee and co-workers Citation[5], Citation[47] evaluated thermal selection responses in TRPV4 mice. In thermal gradient and thermal selection assays, TRPV4-knockout mice preferred higher temperatures than wild type controls. In tail immersion tests, withdrawal responses were prolonged at 45 and 46°C, but were not different from wild type animals between 47 and 50°C. Inconsistencies between tail immersion Citation[5], Citation[47] and hot plate results Citation[37], Citation[41], Citation[42] may be attributed to differences in sensitivities of the tail and paws, how the tests were scored, or in genetic background. In the hyperalgesia assay with complete Freund's adjuvant, no genotype differences were observed in the thermal selection assay Citation[5], Citation[47]. These findings are in contrast to those from Todaka and colleagues Citation[42] and bring into question the role of TRPV4 in hyperalgesia. In a separate experiment, body temperatures were not different between genotypes in terms of circadian rhythmicity at 25°C or following an abrupt increase in ambient temperature from 25 to 35°C Citation[5], Citation[47]. These findings support the idea that TRPV4 is not required for thermoregulation, but is necessary for detection of thermal changes in the periphery.

Aside from nervous tissues, TRPV4 channels have been found in kidney, endothelial cells, and keratinocytes. In kidney, TRPV4 is postulated to play a role in systemic osmoregulation Citation[39]. In endothelium deletion of the TRPV4 gene has been found to alter membrane permeability and Ca2+-influx Citation[40], Citation[48–52]. Keratinocytes have been found to contain several different types of TRP channels, in particular they possess TRPV3 and TRPV4 channels Citation[30], Citation[32], Citation[36]. Although TRPV1 channels have been found in human keratinocytes Citation[53], Citation[54], they have not been reliably identified in mouse keratinocytes Citation[33]. However, whether all three channels exist within the same cells and, if co-expressed, how they may function within the same cell is unknown. Chung and colleagues Citation[33] conducted an electrophysiology experiment with primary keratinocytes to address this question for TPRV3 and TRPV4 channels. They found these cells detect warm temperatures by two distinct mechanisms. The TRPV3 responses were characterized by a slow increase in current amplitude and a protracted sensitization to repetitive warm thermal stimulation. The TRPV4 responses appeared to involve rapidly activating, weakly rectifying, desensitizing currents. The latter responses were absent in TRPV4-knockout cells and they re-appeared when the null cells were transfected with murine TRPV4 cDNA. Hence, keratinocytes possess distinct responses mediated by TRPV3 and TRPV4 channels to signify changes in mild temperatures. Additional research has shown that TRPV3 and TRPV4 form homodimers Citation[35]. Presently, it is unclear whether they can form heterodimers with hybrid functional characteristics.

Keratinocytes are found primarily in the outer layer of the skin where they form tight junctions with nerves of the skin, and can modulate lymphocyte and Langerhans cell function Citation[55]. Keratinocytes are involved in making keratin that serves to protect the skin and underlying tissue from environmental damage (e.g., UV radiation, heat, and water loss). These cells arise from proliferative cells located in the basal layer of the epidermis and migrate to the surface. Additional epithelial cells include melanocytes, Langerhans cells, Merkel cells, and sensory nerve terminals. The inner skin layer or dermis contains connective tissue, hair follicles, sebaceous and sweat glands, nerves and sympathetic fibers, blood and lymphatic vessels, and specialized cells such as Pacinian corpuscles. Results from the TRPV3 and some TRPV4 experiments indicate that keratinocytes participate in thermal sensation. What is not clear is how keratinocytes can communicate with sensory neurons. It has been proposed that the non-neuronal keratinocytes may signal to sensory neurons through the release of some soluble factor (perhaps ATP) Citation[5], Citation[56]. This type of mechanism has already been reported for the gustatory system for tastant detection by epithelial taste buds and gustatory sensory afferent terminals Citation[57]. Since non-peptidergic and peptidergic nociceptive innervations of epithelia are segregated spatially within the skin and spinal cord Citation[58], it may be the case that similar parallel distinct neural pathways exist for sending temperature information from keratinocytes to sensory neurons.

TRP channels that sense cold

Although TRPV1, TRPV3, and TRPV4 play roles in heat detection, TRPM8 (TRP channel, subfamily M, member 8) and TRPA1 (TRP channel, subfamily A, member 1) or ANKTM1 (ankyrin-repeat, transmembrane, member 1) are responsive to cool or cold temperatures. Typically, innocuous cool includes temperatures from 30 to 15°C with noxious cold involving temperatures below 15°C Citation[59–61]. TRPM8 channels respond to “cooling” compounds such as menthol and to reductions in temperature Citation[62–64]. It should be emphasized that TRPM8 demonstrates steep temperature-dependence that extends from cool (<28°C) to noxious cold (<15°C) Citation[65] (). Interestingly, reductions in intracellular pH inhibit the cold-induced activation of TRPM8 Citation[66]. Since in whole-cell recordings of excised patches TRPM8 can be activated by changes in temperature, it is likely that this receptor is stimulated directly by cool temperatures and that it does not require some soluble factor for activation Citation[67]. TRPM8 is expressed in 5–10% of the small diameter tyrosine kinase receptor (TrkA) positive cells in dorsal root ganglion – suggestive of C-fiber and Aδ-fiber localizations; however, these cells do not co-express calcitonin gene-related peptide or substance P or other markers of nociception Citation[63]. In a more recent study, investigators targeted green fluorescent protein to TRPM8-expressing cells using bacterial artificial chromosome technology and demonstrates that this channel is expressed in C-fibers and Aδ-fibers of the trigeminal or dorsal root ganglion Citation[68]. These neurons terminate as peripheral nerve endings in skin in areas that mediate responses to innocuous cool, noxious cold, and nociception.

The gene for Trpm8 has been disrupted in mice by three independent groups Citation[69–71]. Bautista and colleagues Citation[69] observed trigeminal neurons from TRPM8-knockout mice to be unresponsive to menthol or icilin, while Colburn and coworkers Citation[70] found menthol sensitivity to be reduced in their mutants. Menthol responses were reduced also in mutant dorsal root ganglia neurons Citation[71]. When cold-sensitive trigeminal neurons were subjected to a cooling gradient (30 to 8°C over 30 seconds) or to a 18°C stimulus, responses were distributed according to menthol-sensitive and –insensitive neurons Citation[69]. The former neurons had an average threshold of 22°C, but when cooled from a holding temperature of 22°C the activation threshold shifted to 16°C – suggesting plasticity of response and adaptation. The latter neurons had a slower cold activation rate and had an average response threshold of 12°C regardless of starting temperature. Importantly, they were not mustard oil sensitive suggesting that TRPM8 does not participate in this response Citation[69]. Deletion of TRPM8 only affected the menthol-sensitive group where the numbers of cells and magnitude of responses were diminished Citation[69]. Despite this fact, TRPM8-deficient neurons displayed normal responses to heat. Similar results were seen with skin-nerve preparations where responses in C-fibers and Aδ-fibers were reduced in TRPM8-knockout samples, while mechanical responses were intact Citation[69]. In behavioral tests mutant mice were deficient in the acetone-evoked evaporative cooling test Citation[69], Citation[71]. Mutants also could not discriminate warmer from cooler surfaces, but could detect cold – however, the thresholds for cold detection were different among the groups Citation[69–71]. No genotype differences were observed for detection of heat or pressure. It should be noted that the differences in cold detection thresholds were not unexpected because of voltage-dependent gating characteristics of TRPM8 (as well as TRPV1) render these channels sensitive to temperature-induced shifts in their voltage-dependent activation curves Citation[56], Citation[64], Citation[72]. When cold-plate responses were examined at 0°C, wild type mice responded within ∼8 seconds, whereas TRPM8-knockouts took approximately twice as long but their individual latencies were still much shorter than the maximal 90 seconds allotted for the test Citation[70]. However, Dhaka and colleagues Citation[71] failed to observe a genotype difference at −1°C – perhaps because the temperature-dependence of the cold plate response is so steep. Although responses to mechanical allodynia were intact, neuropathic cold responses were deficient in mutant mice Citation[70]. TRPM8-knockout mice were evaluated also whether cold could be analgesic to formalin injection Citation[71]. On a 17°C cold plate, formalin-treated wild type mice displayed reduced responses during the acute (0–10 minutes after injection) and inflammatory (10–30 minutes after injection) phases of the test – demonstrating that cold is analgesic to formalin administration. TRPM8-knockout mice did not display the cold-induced analgesic response during the first 10 minutes of the test. Collectively, these data suggest that TRPM8 is important for thermal sensitivity and that it contributes to responses in the cold range; however, detection of noxious cold appears to involve additional mechanisms that may be due to other receptors, changes in vascular tone, alterations in nociception, or indirect effects from tissue damage or inhibition of warm-sensing fibers. The temperature and nociceptive results suggest that TRPM8 may be expressed at least in two different sets of neurons. One responds to cool temperatures to affect the perception of cool stimuli to control behavioral thermoregulation and analgesia, and a second population transmits nociceptive information to influence nocifensive behaviors Citation[8].

ANKTM1 or TRPA1, another member of the TRP family, has been cloned Citation[73]. TRPA1 responds to cold temperature (<18°C), cannabinoids, mustard oil and other isothiocyanates, as well as mechanical stimuli (hair cells in inner ear), but not menthol Citation[56], Citation[74] (). TRPA1 is expressed in a small population of TRPV1-containing, but not TRPM8-containing, sensory neurons, as well as with the nociceptive markers calcitonin gene-related peptide and substance P neurons Citation[56], Citation[75]. TRPA1 is not expressed with TRPM8 – indicating these channels do not coordinate at the level of individual sensory cells to integrate thermal information. Additionally, since TRPA1 is not expressed in heavily myelinated neurons, it has been suggested that it is localized to non-myelinated C- or lightly myelinated Aδ-fibers. Different splice forms of TRPA1 have been reported Citation[56]. The short form does not posses transmembrane domains and does not constitute a functional channel. It serves to block transport of the long form of the receptor from the cytoplasm to the membrane. Experiments in heterologous cells have shown that TRPA1 is stimulated by formalin, whereas other members of the TRP family are not responsive Citation[76]. Neurons from wild type trigeminal and dorsal root ganglia were activated also by formalin and the responses were blocked by a TRPA1 inhibitor (HC-030031), whereas neurons from TRPA1-knockout mice were unresponsive to formalin. Similarly, HC-030031 also blocked formalin-induced paw lifting, licking and flinching (Phase I), as well as, sensitization responses (Phase II) in wild type mice, but all formalin-induced behaviors were reduced in the knockout animals. Hence, TRPA1 appears to mediate nociceptive responses to formalin.

The temperature sensitivity of TRPA1 has been questioned by some investigators. For instance, when evaluating neurons that responded to cold (5°C), 95% were activated by menthol and the remainder were insensitive to mustard oil and menthol Citation[75]. Only ∼4% of the trigeminal neurons that responded to mustard oil were also activated by cold stimulus (5°C), but all of the cold-sensitive neurons were stimulated by menthol – suggesting that they had TRPM8 receptors Citation[75]. The gene for Trpa1 has been disrupted by two independent groups Citation[77], Citation[78]. Neurons from TRPA1-knockout mice were insensitive to allicin and mustard oil, but were fully activated by capsaicin, menthol, and Ca2+ Citation[77]. When mustard oil was applied to the hind-paw, licking and flinching were abrogated in the knockout animals and inflammation was attenuated. Additionally, TRPA1-knockout mice showed no evidence of hypersensitivity to radiant heat or mechanical stimulation. When sensitivity of trigeminal neurons to cold stimuli (bath was reduced from ∼23° to 6°C) was examined, 78% of the wild type cold-sensitive neurons responded to menthol while the remaining neurons were insensitive to it or mustard oil. A similar distribution was observed in TRPA1-knockout samples and these results were replicated in dorsal root ganglia neurons. In behavior no genotype differences were observed for acute nocifensive responses of shivering or paw withdrawal to cold surfaces (20, 10, 5, 0, −5, or −10°C), preference of cold surfaces over those at room temperature, acetone-evoked evaporative cooling, or for vestibular or auditory responses. Bradykinin has long been known to exert two actions on nociceptors: acute depolarization and protracted hypersensitivity to mechanical or thermal stimuli Citation[79], Citation[80]. When examined for phospholipase C-mediated bradykinin responses, trigeminal neurons from TRPA1-knockout mice were activated less by bradykinin than those from wild type animals Citation[77]. When bradykinin was injected into the hind-paws of knockout mice and they were exposed to radiant heat, no evidence for hypersensitivity was observed. Nevertheless, both wild type and mutant mice developed thermal hypersensitivity to Freund's complete adjuvant which stimulates multiple inflammatory pathways. These findings suggest that TRPA1-knockout mice are responsive to a subset of inflammatory agents that activate phospholipase C signaling pathways. Results from these studies suggest that mustard oil and allicin stimulate TRPA1 receptors on primary afferent neurons to produce inflammatory pain, but do not mediate responses to cold temperatures or vestibular or auditory stimuli.

By comparison, some investigators find support for Story and colleagues’ Citation[73] contention that TRPA1 channels are activated by cold. Kwan and co-workers Citation[78] have also disrupted Trpa1 in mice. In concert with Bautista and collaborators Citation[77], their mutants were not deficient in vestibular or auditory function and electrophysiological responses to mustard oil in neurons for mutant mice were almost completely lost. In a behavioral test, mustard oil was added to the drinking water and at the highest concentration wild type and heterozygous animals ceased drinking; homozygous mutants consumed a third less than normal. TRPA1-knockout mice were also less responsive than wild type animals to mustard oil injection in their hind-paw. Additionally, mutants were also less sensitive than wild type controls to bradykinin-induced mechanical hyperalgesia and to bradykinin injection into the hind-paw. Although both genotypes showed similar responses to a hot-plate (50, 52, 55°C), paw withdrawal responses to the cold plate (0°C) were reduced in mutants Citation[78]. TRPA1-knockout mice were also less sensitive to acetone-induced evaporative cooling than wild type animals. Interestingly, the genotype difference was larger for females than males on the cold plate and acetone tests. Additionally, knockout mice showed reduced mechanical sensitivity to von Frey filaments and to blunt pressure in the Randall-Selitto test. In a model of peripheral neuropathic pain (ligation of the tibial and common peroneal nerves), no genotype differences in mechanical hypersensitivity were noted. Since this response is thought to be mediated by low-threshold, large, myelinated non-peptidergic Aβ-fibers Citation[80], Citation[81] and because TRPA1 is not expressed in these fibers Citation[82], Citation[83], the lack of a genotype distinction is consistent with TRPA1 expression results. Additional convincing evidence for TRPA1 playing some role in thermal sensitivity comes from work by Obata and co-workers Citation[84]. Here, pharmacological blockade of TRPA1 in primary sensory neurons was sufficient to reverse the cold hyperalgesia produced by nerve injury and peripheral inflammation. Moreover, both conditions increased expression of TRPA1, but not TRPM8, in TrkA-containing dorsal root neurons. Intrathecal injection of TRPA1 antisense oligonucleotides depressed the induction of TRPA1 and cold hyperalgesia. More recent experiments using cultured vagal sensory neurons from nodose ganglion have identified a large fraction of neurons that respond to cold with a threshold ∼24°C Citation[85]. The responses of these neurons were evaluated with TRPM8 and TRPA1 agonists and antagonists and these experiments suggested that TRPA1 channels were the primary mediators of the cold responses. These results were replicated in mouse nodose ganglion and neurons from TRPA1-knockout mice [note, mice from Citation[78] showed a large reduction in the percentage of cold-responsive neurons. Similarly, Karashima and colleagues Citation[86] found TRPA1 in heterologous cells to be responsive to the cold (26 to 10°C) and for trigeminal mustard oil responsive neurons from TRPA1-knockout mice [mice from Citation[78] to be deficient in cold sensitivity. In wild type mice, TRPA1 trigeminal neurons had a lower threshold for activation to cool and cold temperatures, showed different time-courses, and were differentially sensitive to capsaicin than TRPM8 neurons. On the cold plate (0°C) and cold tail-flick (−10°C) tests, TRPA1-knockout mice were deficient compared to their wild type controls Citation[86]. Hence, the bulk of the evidence supports a role for TRPA1 in mediating effects of cold temperatures. There may be several reasons for disagreement in the investigation of TRPA1-knockout mice. First, a fragment of TRPA1 was still expressed in Bautista and colleagues’ Citation[77] knockout animals. While this fragment did not appear to function as a dominant negative for capsaicin, menthol sensitivity, or Ca2+ handling, it may have exerted some unrealized effect on cold responses. Second, the presumed TRPA1-mediated behavioral responses to isothiocyanates could have been modulated by other receptors. Third, because various chemical agonists and thermal stimuli may gate the TRPA1 channel and activate other nociceptors, the final psychophysical responses may represent a summation or some combination of these inputs. In an analogous fashion responses to noxious cold may require additional inputs besides TRPA1 from converging neurons containing other receptors sensing cool (e.g., TRPM8). Finally, there is evidence that some cold-responsive neurons in dorsal root and superior cervical ganglia do not respond to menthol or mustard oil Citation[87]. The threshold for cold temperatures was higher in dorsal root ganglion where only TRPM8 was expressed; TRPA1 was expressed at 10-fold higher levels in superior cervical ganglion than dorsal root ganglion. Hence, there may be additional receptors besides TRPM8 and TRPA1 that are sensitive also to cold. It this respect, Kv1 potassium channels can modulate the threshold for cold stimuli in TRPM8-expressing cells Citation[88]. Additional non-TRP candidates for temperature regulation include background or TREK-1 potassium channels Citation[67], Citation[89], Citation[90], Na+/K+ ATPase Citation[91], the DEG/ENaC sodium channels Citation[92], the voltage-gated sodium channel NaV1.8 Citation[93], Citation[94], and the ATP-gated P2X3 purinergic receptor Citation[91].

Heteroreceptor regulation of TRP channels

Activities of many TRP family members can be modulated by G protein coupled and Trk receptors that are coupled to phospholipase C and production of diacylglycerol, inositol phosphates, and/or changes in extracellular or intracellular Ca2+ concentrations Citation[4], Citation[91]. Parenthetically, there is evidence that some TRP channels can be modulated also by eicosanoids, endocannabinoids, and various other signaling molecules Citation[4], Citation[56], Citation[91], Citation[95], Citation[96]. TRPV1 channels can be heterologously sensitized through actions of G protein coupled or tyrosine kinase receptors so that the channel opens at normal body temperature Citation[97–99]. Many other TRP family members also show similar responses; however, activation of the TRPA1 channel can occur in the absence of Ca2+ and with Ca2+ availability augmenting the response Citation[85]. Additionally, the TRPA1 channel has been proposed to be activated by bradykinin in two ways: through activation of the bradykinin receptor and stimulation of phospholipase C-mediated increases in intracellular Ca2+ and by Ca2+ influx through TRPV1 channels Citation[77]. Since TRPA1 can be activated by thermal, mechanical, and various chemical stimuli, it may be an important integrator of information. Since other TRP channels can also be polymodal transducers, their responses to thermal, nociceptive, chemical, and other changes may show plasticity where the threshold for activation and dose-response curve are shifted.

Promiscuous TRP channels

Although TRPV1, TRPV3, TRPV4, TRPM8, and TRPA1 channels respond to temperature, many also respond to pain. For instance, TRPV1 channels respond to inflammatory pain Citation[16] and this condition results in peripheral sensitization where the threshold for activation is decreased. Under these circumstances, the threshold for heat pain is reduced from ∼42°C to a level that is near body temperature Citation[56]. In a model of peripheral nerve injury, TRPV1 levels are reduced in injured neurons but they increase in neighboring neurons Citation[100]; thereby, potentially altering sensitivity to thermal stimuli. TRPV1 channels can be rendered more sensitive to stimuli through phosphorylation of the channel Citation[101–103]; they can also be desensitized by dephosphorylation Citation[104–105]. It should be noted, that local application of capsaicin has been used therapeutically for many years to produce analgesia for treatment of postherpetic neuralgia and diabetic neuropathy Citation[106]. This therapy also reduces the sensitivity of the skin to thermal agents Citation[107].

Development of compounds that target TRP channels

Although TRPV1 is the most studied and validated thermally-sensitive TRP channel, development of agonists and antagonists for this and other TRP family members are currently undergoing intense development Citation[11], Citation[22], Citation[108]. Three different strategies have been applied: development of receptor agonists and receptor antagonists, as well as, developing compounds that can enter the pore of the TRP channels and exert actions. For instance, TRPV1 channels show time- and agonist-dependent increases in permeability that is due to phosphorylation of the channel by protein kinase C Citation[109]. QX-314, a lidocane derivative, cannot block Na+ channels when it is administered extracellularly because it cannot gain access to the inner face of the Na+ channel Citation[110]. When administered with capsaicin, QX-314 can enter TRPV1-containing cells through the TRPV1 channel and exert effects that last more than 10-times longer than lidocane alone. Notably, the threshold for pain and radiant heat sensitivities are reduced by this agent. In comparison to this mechanism, chronic administration of TRPV1 agonists can be used to desensitize the receptor, such that its actions are similar to that of antagonists that block actions at the receptor.

As described above, the TRPV1 channel and some other TRP family members may be activated by many stimuli. In the case of TRPV1, different domains of the receptor have been found to be responsive to different stimuli. For instance, the C-terminal region confers the channel with heat sensitivity Citation[111]. Some investigators have been able to develop TRPV1 compounds (i.e., AMG-8562) that can antagonize responses to capsaicin and acid, but leave the heat-evoked responses intact Citation[112]. This indicates that it is possible to target different functional domains of the receptor for selective activities. However, it should be emphasized that the pharmacological effects of antagonists do not always mimic results reported in the TRPV1-knockout mice Citation[11], Citation[21]. These differences may be attributed to off-target effects of the compounds, alterations in sensitization and/or down-regulation of the channels in different tissues with subchronic or chronic dosing, and compensation of some functions in the knockout mice. Additionally, long-term effects of administration have not been investigated. The future challenge in developing compounds that target TRP channels will be to stimulate, attenuate, or block certain physiological responses while leaving others intact.

TRP channels and studies of hyperthermia

The cloning and identification of receptors that can sense thermal changes present a number of new avenues for investigation for hyperthermia research. Aside from localization to sensory neurons and skin, TRP channels have been identified in other tissues. Currently, their physiological roles in these other tissues are unknown. It is also unclear the extent to which normal physiological responses to hyperthermia in brain and other tissues are mediated by these receptors. For instance, increases in body temperature are well-known to affect cognitive function. Agonist stimulation of TRPV1 channels has been reported to block long-term potentiation in hippocampus Citation[10]. In this regard, it will be important in the future to determine whether hyperthermic conditions activate TRP channels in brain and other tissues and lead to changes in local tissue functions and behavior. Finally, one may anticipate that over time, polymorphisms in various TRP channel genes will be identified in humans and other organisms. In this context, it will be important to ascertain which mutations may render the individual more or less sensitive to hyperthermic stimuli or whether the physiological changes ascribed to hyperthermia are mediated by other mechanisms.

Acknowledgements

The contents of this review were supported by Microwave Manufacturers Foundation and the GSM Association.

Declaration of interest: I wish to declare that I have no conflicts of interests in this work.

References

  • Montell C, Rubin GM. Molecular characterization of the Drosophila trp locus: A putative integral membrane protein required for phototransduction. Neuron 1989; 2: 1313–1323
  • 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
  • Caterina MJ, Rosen TA, Tominaga M, Brake AJ, Julius D. A capsaicin-receptor homologue with a high threshold for noxious heat. Nature 1999; 398: 436–441
  • Ramsey IS, Delling M, Clapham DE. An introduction to TRP channels. Annu Rev Physiol 2006; 68: 619–647
  • Lee H, Caterina MJ. TRPV channels as thermosensory receptors in epithelial cells. Pflugers Arch 2005; 451: 160–167
  • Tominaga M, Caterina MJ, Malmberg AB, Rosen TA, Gilbert H, Skinner K, Raumann BE, Basbaum AI, Julius D. The cloned capsaicin receptor integrates multiple pain-producing stimuli. Neuron 1998; 21: 531–543
  • Jordt S-E, Julius D. Molecular basis for species-specific sensitivity to hot chili peppers. Cell 2002; 108: 421–430
  • Steiner AA, Turek VF, Almeida MC, Burmeister JJ, Oliveira DL, Roberts JL, Bannon AW, Norman MH, Louis J-C, Treanor JJS, 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
  • Chung M-K, Caterina MJ. TRP channel knockout mice lose their cool. Neuron 2007; 54: 345–347
  • Kauer JA, Gibson HE. Hot flash: TRPV channels in the brain. Trends Neurosci 2009; 32: 215–224
  • Szallasi A, Cortright DN, Blum CA, Eid SR. The vanilloid receptor TRPV1: 10 years from channel cloning to antagonist proof-of-concept. Nat Rev Drug Discov 2007; 6: 2357–372
  • Wang C, Hu HZ, Colton CK, Wood JD, Zhu MX. An alternative splicing product of the murine trpv1 gene dominant negatively modulates the activity of TRPV1 channels. J Biol Chem 2004; 279: 37423–37430
  • Lu G, Henderson D, Liu L, Reinhart PH, Simon SA. TRPV1b: A functional human vanilloid receptor splice variant. Mol Pharmacol 2005; 67: 1119–1127
  • Sharif Naeini R, Witty MF, Seguela P, Borque CW. An N-terminal variant of Trpv1 channel is required for osmosensory transduction. Nature Neurosci 2006; 9: 93–98
  • Tian W, Wang DH, Cohen DM. Regulation of TRPV1 by a novel renally expressed TRPV1 splice variant. Am J Physiol Renal Physiol 2006; 290: F117–F126
  • Caterina MJ, Leffler A, Malmberg AB, Martin WJ, Trafton J, Petersen-Zeitz KR, Koltzenburg M, Basbaum AI, Julius D. Impaired nociception and pain sensation in mice lacking the capsaicin receptor. Science 2000; 288: 306–313
  • Davis JB, Gray J, Gunthorpe MJ, Hatcher JP, Davey PT, Overend P, Harries MH, Latcham J, Clapham C, Atkinson K, et al. Vanilloid receptor-1 is essential for inflammatory thermal hyperalgesia. Nature 2000; 405: 183–187
  • Iida T, Shimizu I, Nealen ML, Campbell A, Caterina M. Attenuated fever response in mice lacking TRPV1. Neurosci Lett 2005; 378: 28–33
  • Hori T. Capsaicin and central control of thermoregulation. Pharmacol Ther 1984; 26: 389–416
  • Szallasi A, Blumberg PM. Vanilloid (capsaicin) receptors and mechanisms. Pharmacol Rev 1999; 51: 159–212
  • Szelényi Z, Hummel Z, Szolcsányi J, Davis JB. Daily body temperature rhythm and heat tolerance in TRPV1 knockout and capsaicin pretreated mice. Eur J Neurosci 2004; 19: 1421–1424
  • Gavva NR. Body-temperature maintenance as the predominant function of the vanilloid receptor TRPV1. Trends Pharmacol Sci 2008; 29: 550–557
  • Gavva NR, Bannon AW, Surapaneni S, Hovland DN, Jr, Lehto SG, Gore A, Juan T, Deng H, Han B, Klionsky L, et al. The vanilloid receptor TRPV1 is tonically activated in vivo and involved in body temperature regulation. J Neurosci 2007; 27: 3366–3374
  • Gavva NR, Treanor JJS, Garami A, Fang L, Surapaneni S, Akrami A, Alvarez F, Bak A, Darling M, Gore A, et al. Pharmacological blockade of the vanilloid receptor TRPV1 elicits marked hyperthermia in humans. Pain 2008; 136: 202–210
  • Caterina MJ. Transient receptor potential ion channels as participants in thermosensation and thermoregulation. Am J Physiol Regul Integr Comp Physiol 2007; 292: R64–R76
  • Jordt S-E, McKemy DD, Julius D. Lessons from peppers and peppermint: The molecular logic of thermosensation. Curr Opin Neurobiol 2003; 13: 487–492
  • Woodbury CJ, Zwick M, Wang S, Lawson JJ, Caterina MJ, Koltzenburg M, Albers KM, Koerber HR, Davis BM. Nociceptors lacking TRPV1 and TRPV2 have normal heat responses. J Neurosci 2004; 24: 6410–6415
  • Lawson JJ, McIlwrath SL, Woodbury CJ, Davis BM, Koerber HR. TRPV1 unlike TRPV2 is restricted to a subset of mechanically insensitive cutaneous nociceptors responding to heat. J Pain 2008; 9: 298–308
  • Moqrich A, Hwang SW, Earley TJ, Petrus MJ, Murray AN, Spencer KS, Andahazy M, Story GM, Patapoutian A. Impaired thermosensation in mice lacking TRPV3, a heat and camphor sensor in the skin. Science 2005; 307: 1468–1472
  • Peier AM, Reeve AJ, Andersson DA, Moqrich A, Earley TJ, Hergarden AC, Story GM, Colley S, Hogenesch JB, McIntyre P, et al. A heat-sensitive TRP channel expressed in keratinocytes. Science 2002; 296: 2046–2049
  • Smith GD, Gunthorpe MJ, Kelsell RE, Hayes PD, Reilly P, Facer P, Wright JE, Jermank JC, Walhin J-P, Ooi L, et al. TRPV3 is a temperature-sensitive vanilloid receptor-like protein. Nature 2002; 418: 186–190
  • Xu H, Ramsey IS, Kotecha SA, Moran MM, Chong JA, Lawson D, Ge P, Lilly J, Silos-Santiago I, Xie Y, et al. TRPV3 is a calcium-permeable temperature-sensitive cation channel. Nature 2002; 418: 181–186
  • Chung M-K, Lee H, Mizuno A, Suzuki M, Caterina MJ. TRPV3 and TRPV4 mediate warmth-evoked currents in primary mouse keratinocytes. J Biol Chem 2004; 279: 21569–21575
  • Chung M-K, Güler AD, Caterina MJ. Biphasic currents evoked by chemical or thermal activation of the heat-gated ion channel. J Biol Chem 2005; 280: 15928–15941
  • Hellwig N, Albrecht N, Harteneck C, Schultz G, Schaefer M. Homo- and heteromeric assembly of TRPV channel subunits. J Cell Sci 2005; 118: 917–928
  • 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
  • Liedtke W, Friedman JM. Abnormal osmotic regulation in trpv4-/- mice. Proc Natl Acad Sci USA 2003; 100: 13698–13703
  • Mizuno A, Matsumoto N, Imai M, Suzuki M. Impaired osmotic sensation in mice lacking TRPV4. Am J Physiol 2003; 285: C96–C101
  • Cohen DM. TRPV4 channels in the mammalian kidney. Pflugers Archiv Eur J Physiol 2005; 451: 168–75
  • Hartmannsgruber V, Heyken WT, Kacik M, Kaistha A, Grgic I, Harteneck C, Liedtke W, Hoyer J, Köhler R. Arterial response to shear stress critically depends on endothelial TRPV4 expression. PLoS One 2007; 2(9)e827
  • Suzuki M, Mizuno A, Kodaira K, Imai M. Impaired pressure sensation in mice lacking TRPV4. J Biol Chem 2003; 278: 22664–22668
  • Todaka H, Taniguchi J, Satoh J, Mizuno A, Suzuki M. Warm temperature-sensitive transient receptor potential vanilloid 4 (TRPV4) plays an essential role in thermal hyperalgesia. J Biol Chem 2004; 279: 35133–35138
  • Hargreaves K, Dubner R, Brown F, Flores C, Joris J. A new and sensitive method for measuring thermal nociception in cutaneous hyperalgesia. Pain 1988; 32: 77–88
  • Vriens J, Watanabe H, Janssens A, Droogmans G, Voets T, Nilius B. Cell swelling, heat, and chemical agonists use distinct pathways for the activation of the cation channel TRPV4. Proc Natl Acad Sci USA 2004; 101: 396–401
  • Watanabe H, Vriens J, Prenen J, Droogmans G, Voets T, Nilius B. Anandamide and arachidonic acid use epoxyeicosatrienoic acids to activate TRPV4 channels. Nature 2003; 424: 434–438
  • Watanabe H, Vriens J, Suh SH, Benham CD, Droogmans G, Nilius B. Heat-evoked activation of TRPV4 channels in a HEK293 cell expression system and in native mouse aorta endothelial cells. J Biol Chem 2002; 277: 47044–47051
  • Lee H, Iida T, Mizuno A, Suzuki M, Caterina MJ. Altered thermal selection behavior in mice lacking transient receptor potential vanilloid 4. J Neurosci 2005; 25: 1304–1310
  • Reiter B, Kraft R, Günzel D, Zeissig S, Schulzke JD, Fromm M, Harteneck C. TRPV4-mediated regulation of epithelial permeability. FASEB J 2006; 20: 1802–1812
  • Nilius B, Droogmans G, Wondergem R. Transient receptor potential channels in endothelium: Solving the calcium entry puzzle?. Endothelium 2003; 10: 5–15
  • Alvarez DF, King JA, Weber D, Addison E, Liedtke W, Townsley MI. Transient receptor potential vanilloid 4-mediated disruption of the alveolar septal barrier: A novel mechanism of acute lung injury. Circ Res 2006; 99: 988–995
  • Earley S, Pauyo T, Drapp R, Tavares MJ, Liedtke W, Brayden JE. TRPV4-dependent dilation of peripheral resistance arteries influences arterial pressure. Am J Physiol Heart Circ Physiol 2009; 297: H1096–H1102
  • Zhang DX, Mendoza SA, Bubolz AH, Mizuno A, Ge ZD, Li R, Warltier DC, Suzuki M, Gutterman DD. Transient receptor potential vanilloid type 4-deficient mice exhibit impaired endothelium-dependent relaxation induced by acetylcholine in vitro and in vivo. Hypertension 2009; 53: 532–538
  • Denda M, Fuziwara S, Inoue K, Denda S, Akamatsu H, Tomitaka A, Matsunaga K. Immunoreactivity of VR1 on epidermal keratinocyte of human skin. Biochem Biophys Res Commun 2001; 285: 1250–1252
  • Inoue K, Koizumi S, Fuziwara S, Denda S, Denda M. Functional vanilloid receptors in cultured normal human epidermal keratinocytes. Biochem Biophys Res Commun 2002; 291: 124–129
  • Bloom W, Fawcett DW. A textbook of histology. W.B. Saunders Company, Philadelphia 1975; 563–597
  • Wang H, Woolf CJ. Pain TRPs. Neuron 2005; 46: 9–12
  • Kinnamon SC, Cummings TA. Chemosensory transduction mechanisms in taste. Annu Rev Physiol 1992; 54: 715–731
  • Zylka MJ, Rice FL, Anderson DJ. Topographically distinct epidermal nociceptive circuits revealed by axonal tracers targeted to Mrgprd. Neuron 2005; 45: 17–25
  • Hensel H. Thermoreception and temperature regulation. Monogr Physiol Soc 1981; 38: 1–321
  • Spray DC. Cutaneous temperature receptors. Annu Rev Physiol 1986; 48: 625–638
  • Davis KD, Pope GE. Noxious cold evokes multiple sensations with distinct time courses. Pain 2002; 98: 179–185
  • McKemy DD, Neuhausser WM, Julius D. Identification of a cold receptor reveals a general role for TRP channels in thermosensation. Nature 2002; 416: 52–58
  • Peier AM, Moqrich A, Hergarden AC, Reeve AJ, Andersson DA, Story GM, Earley TJ, Dragoni I, McIntyre P, Bevan S, et al. A TRP channel that senses cold stimuli and menthol. Cell 2002; 108: 705–715
  • Voets T, Droogmans G, Wissenbach U, Janssens A, Flockerzi V, Nilius B. The principle of temperature-dependent gating in cold- and heat-sensitive TRP channels. Nature 2004; 430: 748–754
  • McKemy DD. How cold is it? TRPM8 and TRPA1 in the molecular logic of cold sensation. Mol Pain 2005; 1: 16
  • Andersson DA, Chase HW, Bevan S. TRPM8 activation by menthol, icilin, and cold is differentially modulated by intracellular pH. J Neurosci 2004; 24: 5364–5369
  • Reid G, Flonta ML. Ion channels activated by cold and menthol in cultured rat dorsal root ganglion neurones. Neurosci Lett 2002; 324: 164–168
  • Takashima Y, Daniels RL, Knowlton W, Teng J, Liman ER, McKemy DD. Diversity in the neural circuitry of cold sensing revealed by genetic axonal labeling of transient receptor potential melastatin 8 neurons. J Neurosci 2007; 27: 14147–14157
  • Bautista DM, Siemens J, Glazer JM, Tsuruda PR, Basbaum AI, Stucky CL, Jordt SE, Julius D. The menthol receptor TRPM8 is the principal detector of environmental cold. Nature 2007; 448: 204–209
  • Colburn RW, Lubin ML, Stone DJ, Jr, Wang Y, Lawrence D, D'Andrea MR, Brandt MR, Liu Y, Flores CM, Qin N. Attenuated cold sensitivity in TRPM8 null mice. Neuron 2007; 54: 379–386
  • Dhaka A, Murray AN, Mathur J, Earley TJ, Petrus MJ, Patapoutian A. TRPM8 is required for cold sensation in mice. Neuron 2007; 54: 371–378
  • Mälkiä A, Madrid R, Meseguer V, de la Peña E, Valero M, Belmonte C, Viana F. Bidirectional shifts of TRPM8 channel gating by temperature and chemical agents modulate the cold sensitivity of mammalian thermoreceptors. J Physiol 2007; 581: 155–174
  • Story GM, Peier AM, Reeve AJ, Eid SR, Mosbacher J, Hricik TR, Earley TJ, Hergarden AC, Andersson DA, Hwang SW, et al. ANKTM1, a TRP-like channel expressed in nociceptive neurons, is activated by cold temperatures. Cell 2003; 112: 819–829
  • Corey DP, Garcia-Añoveros J, Holt JR, Kwan KY, Lin SY, Vollrath MA, Amalfitano A, Cheung EL, Derfler BH, Duggan A, et al. TRPA1 is a candidate for the mechanosensitive transduction channel of vertebrate hair cells. Nature 2004; 432: 723–730
  • Jordt S-E, Bautista DM, Chuang H-H, McKemy DD, Zygmunt PM, Högestätt ED, Meng ID, Julius D. Mustard oils and cannabinoids excite sensory nerve fibres through the TRP channel ANKTM1. Nature 2004; 427: 260–265
  • McNamara CR, Mandel-Brehm J, Bautista DM, Siemens J, Deranian KL, Zhao M, Hayward NJ, Chong JA, Julius D, Moran MM, et al. TRPA1 mediates formalin-induced pain. Proc Natl Acad Sci USA 2007; 104: 13525–13530
  • Bautista DM, Jordt S-E, Nikai T, Tsuruda PR, Read AJ, Poblete J, Yamoah E, Basbaum AI, Julius D. TRPA1 mediates the inflammatory actions of environmental irritants and proalgesic agents. Cell 2006; 124: 1269–1282
  • Kwan KY, Allchorne AJ, Vollrath MA, Christensen AP, Zhang D-S, Woolf CJ, Corey DP. TRPA1 contributed to cold, mechanical, and chemical nociception but is not essential for hair-cell transduction. Cell 2006; 50: 277–289
  • Dray A, Perkins M. Bradykinin and inflammatory pain. Trends Neurosci 1993; 16: 99–104
  • McMahon SB, Bennett DLH, Bevan S. Inflammatory mediators and modulators of pain. Textbook of pain, SB McMahon, M Koltzenburg. Elsevier, Philadelphia 2006; 49–72
  • Campbell JN, Raja SN, Meyer RA, Mackinnon SE. Myelinated afferents signal the hyperalgesia associated with nerve injury. Pain 1988; 32: 89–94
  • Koltzenburg M, Torebjork HE, Wahren LK. Nociceptor modulated central sensitization causes mechanical hyperalgesia in acute chemogenic and chronic neuropathic pain. Brain 1994; 117: 579–591
  • Kobayashi K, Fukuoka T, Obata K, Yamanaka H, Dai Y, Tokunaga A, Noguchi K. Distinct expression of TRPM8, TRPA1, and TRPV1 mRNAs in rat primary afferent neurons with Aδ/C-fibers and colocalization with Trk receptors. J Comp Neurol 2005; 493: 596–606
  • Obata K, Katsura H, Mizushima T, Yamanaka H, Kobayashi K, Dai Y, Fukuoka T, Tokunaga A, Tominaga M, Noguchi K. TRPA1 induced in sensory neurons contributes to cold hyperalgesia after inflammation and nerve injury. J Clin Invest 2005; 115: 2393–2401
  • Fajardo O, Meseguer V, Belmonte C, Vianna F. TRPA1 channels mediate cold temperature sensing in mammalian vagal sensory neurons: Pharmacological and genetic evidence. J Neurosci 2008; 28: 7863–7875
  • Karashima Y, Talavera K, Everaerts W, Janssens A, Kwan KY, Vennekens R, Nilius B, Voets T. TRPA1 acts as a cold sensor in vitro and in vivo. Proc Natl Acad Sci USA 2009; 106: 1273–1278
  • Munns C, Al Qatari M, Koltzenburg M. Many cold sensitive peripheral neurons of the mouse do not express TRPM8 or TRPA1. Cell Calcium 2007; 41: 331–342
  • Madrid R, de la Pena E, Donovan-Rodriguiz T, Belmonte C, Viana F. Variable threshold of trigeminal cold-thermosensitive neurons is determined by a balance between TRPM8 and Kv1 potassium channels. J Neurosci 2009; 29: 3120–3131
  • Viana F, de la Pena E, Belmonte C. Specificity of cold thermotransduction is determined by differential ionic channel expression. Nat Neurosci 2002; 5: 254–260
  • Maingret F, Lauritzen I, Patel AJ, Heurteaux C, Reyes R, Lesage F, Lazdunski M, Honoré E. TREK-1 is a heat-activated background K+ channel. EMBO J 2000; 19: 2483–2491
  • Dhaka A, Viswanath V, Patapoutian A. TRP ion channels and temperature sensation. Annu Rev Neurosci 2006; 29: 135–161
  • Askwith CC, Benson CJ, Welsh MJ, Snyder PM. DEG/ENaC ion channels involved in sensory transduction are modulated by cold temperature. Proc Natl Acad Sci USA 2001; 98: 6459–6463
  • Zimmermann K, Leffler A, Babes A, Cendan CM, Carr RW, Kobayashi J-I, Nau C, Wood JN, Reeh PW. Sensory neuron sodium channel Nav1.8 is essential for pain at low temperatures. Nature 2007; 447: 855–858
  • Abrahamsen B, Zhao J, Asante CO, Cendan CM, Marsh S, Martinez-Barbera JP, Nassar MA, Dickenson AH, Wood JN. The cell and molecular basis of mechanical, cold, and inflammatory pain. Science 2008; 321: 702–705
  • Clapham DE. TRP channels as cellular sensors. Nature 2003; 426: 517–524
  • Woolf CJ, Ma Q. Nociceptors – noxious stimulus detectors. Neuron 2007; 55: 353–364
  • Chuang HH, Prescott ED, Kong H, Shields S, Jordt SE, Basbaum AI, Chao MV, Julius D. Bradykinin and nerve growth factor release the capsaicin receptor from PtdIns(4,5)P2-mediated inhibition. Nature 2001; 411: 957–962
  • Tominaga M, Wada M, Masu M. Potentiation of capsaicin receptor activity by metabotropic ATP receptors as a possible mechanism for ATP-evoked pain and hyperalgesia. Proc Natl Acad Sci USA 2001; 98: 6951–6956
  • Vellani V, Mapplebeck S, Moriondo A, Davis JB, McNaughton PA. Protein kinase C activation potentiates gating of the vanilloid receptor VR1 by capsaicin, protons, heat and anandamide. J Physiol (Lond) 2001; 543: 813–825
  • Hudson LJ, Bevan S, Wotherspoon G, Gentry C, Fox A, Winter J. VR1 protein increases in undamaged DRG neurons after partial nerve injury. Eur J Neurosci 2001; 13: 2105–2114
  • Premkumar LS, Ahern GP. Induction of vanilloid receptor channel activity by protein kinase C. Nature 2000; 408: 985–990
  • Bhave G, Zhu W, Wang H, Brasier DJ, Oxford GS, Gereau RW IV. cAMP-dependent protein kinase regulates desensitization of the capsaicin receptor (VR1) by direct phosphorylation. Neuron 2002; 35: 721–731
  • Numazaki M, Tominaga T, Toyooka H, Tominaga M. Direct phosphorylation of capsaicin receptor VR1 by protein kinase C-ε and identification of two target serine residues. J Biol Chem 2002; 277: 13375–13378
  • Jung J, Shin JS, Lee S-Y, Hwang SW, Koo J, Cho H, Oh U. Phosphorylation of vanilloid receptor 1 by Ca2+/calmodulin-dependent kinase II regulates its vanilloid binding. J Biol Chem 2003; 279: 7048–7054
  • Mohapatra DP, Nau C. Regulation of Ca2+-dependent desensitization in the vanilloid receptor TRPV1 by calcineurin and cAMP-dependent protein kinase. J Biol Chem 2005; 280: 13424–13432
  • Hautkappe M, Roizen MF, Toledano A, Roth S, Jeffries JA, Ostermeier AM. Review of the effectiveness of capsaicin for painful cutaneous disorders and neural dysfunction. Clin J Pain 1998; 14: 97–106
  • Rashid MH, Inoue M, Kondo S, Kawashima T, Bakoshi S, Ueda H. Novel expression of vanilloid receptor 1 on capsaicin-insensitive fibers accounts for the analgesic effect of capsaicin cream in neuropathic pain. J Pharmacol Exp Ther 2003; 304: 940–948
  • Patapoutian A, Tate S, Woolf CJ. Transient receptor potential channels: Targeting pain at the source. Nat Rev Drug Discov 2009; 8: 55–68
  • Chung M-K, Güler AD, Caterina MJ. TRPV1 shows dynamic ionic selectivity during agonist stimulation. Nature Neurosci 2008; 11: 555–564
  • Binshtok AM, Bean BP, Woolf CJ. Inhibition of nociceptors by TRPV1-mediated entry of impermeant sodium channel blockers. Nature 2007; 449: 607–610
  • Brauchi S, Orta G, Salazar M, Rosenmann E, Latorre R. A hot-sensing cold receptor: C-terminal domain determines thermosensation in transient receptor potential channels. J Neurosci 2006; 26: 4835–4840
  • Lehto SG, Tamir R, Deng H, Klionsky L, Kuang R, Le A, Lee D, Louis J-C, Magal E, Manning BH, et al. Antihyperalgesic effects of AMG8562, a novel vanilloid receptor TRPV1 modulator that does not cause hyperthermia in rats. J Pharmacol Exp Ther 2008; 326: 218–229

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.