Skip to main content
Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Nat Neurosci. Author manuscript; available in PMC 2015 Mar 18.
Published in final edited form as:
PMCID: PMC4364402
NIHMSID: NIHMS671116
PMID: 24473265

Why we scratch an itch: the molecules, cells and circuits of itch

Abstract

Itch is described as an irritating sensation that triggers a desire to scratch. However, this definition hardly seems fitting for the millions of people who suffer from intractable itch. Indeed, the Buddhist philosopher Nāgārjuna more aptly stated, “There is pleasure when an itch is scratched. But to be without an itch is more pleasurable still.” Chronic itch is widespread and very difficult to treat. In this review we focus on the molecules, cells and circuits in the peripheral and central nervous systems that drive acute and chronic itch transmission. Understanding the itch circuitry is critical to developing new therapies for this intractable disease.

Itch is defined as an unpleasant sensation that evokes a desire to scratch. In contrast to acute itch that is transient, chronic itch is a persistent, debilitating condition for which there are few treatment options. Chronic itch accompanies a number of skin diseases and systemic conditions, including eczema, kidney failure, liver cirrhosis and some cancers. A variety of neurological disorders also induce severe, chronic itch; for example, multiple sclerosis, diabetic neuropathy and post-herpetic neuralgia (shingles)13. Chronic itch, like chronic pain, can occur without injury or disease, serves no apparent biological purpose and has no recognizable endpoint. Although studies have identified a number of essential molecules in primary afferent neurons that transduce acute itch signals, we are only now beginning to uncover the cellular and molecular players that drive chronic itch in both the peripheral and central nervous system.

Primary afferent sensory neurons detect a multitude of sensory stimuli, including touch, temperature, chronic pain and itch. These neurons have cell bodies in dorsal root ganglia (DRG) and trigeminal ganglia (TG) and project axons to target organs, such as the skin, in the periphery. There is considerable apparent overlap in the neuronal cells and circuits that transmit itch and pain. Although these sensations are easily distinguished, they are intimately linked in a number of other ways. First, pain inhibits itch, as exemplified by the relief one gets from noxious scratching of itchy skin. Second, some analgesics (for example, opioids) can cause itch. Third, some itch-causing compounds (pruritogens) induce inflammatory pain, and, under certain circumstances, some pain-causing compounds (algogens) induce itch. Finally, there is overlap in some of the signaling molecules that transduce itch and pain1. However, whether itch and pain signals are actually transduced by the same populations of DRG neurons remains unknown.

Subsets of primary neurons have been shown to act as sensors for acute and chronic itch stimuli and to promote scratching behaviors. Itch-sensitive neurons are activated by a variety of exogenous pruritogens (itch-causing compounds), as well as pruritogens produced endogenously by epithelial and immune cells (Fig. 1). However, whether the plethora of pruritogens activates the same populations of itch-sensitive neurons remains controversial. Adding to this cellular complexity, interactions among immune cells, sensory neurons and epithelial cells also contribute to both acute and chronic itch1. Indeed, elucidating both the intercellular and intracellular signaling molecules that drive these sensations is a major goal in the field.

An external file that holds a picture, illustration, etc.
Object name is nihms-671116-f0001.jpg

Schematic depicting the cell types involved in the detection of diverse itch stimuli. DRG neurons (blue; left) innervate the skin and can be activated directly by exogenous or endogenous itch-inducing agents (pruritogens) released by keratinocytes, immune cells or neighboring neuronal afferent endings. However, many of these endogenous compounds also activate keratinocytes and various immune cells. In addition, cross-talk between all three cell types, via release of secreted compounds (for example, via neurogenic inflammation), can further modulate cellular responsiveness and itch pathway output. Distinct subsets of sensory neuron afferents innervate the skin and mediate itch signaling (right). The pruritogen histamine activates neurons via the histamine receptor 1 (H1) that leads to the opening of TRPV1 ion channels (red; top right). The pruritogens BAM8–22 and chloroquine activate neurons via MrgprC11 and MgprA3, respectively, leading to opening of TRPA1 channels (red; middle right). The cytokine TSLP activates neurons via TSLPR, which leads to the opening of TRPA1 channels (red; bottom right). The activation of TRPV1 or TRPA1 leads to neuronal depolarization, action potential firing and the transmission of itch signals from the periphery to the CNS. See Table 1 for a complete list of pruritogens and receptors implicated in itch.

In this review we present our views on some of the most exciting advances in itch research over the past few years. We highlight progress in defining some of the molecules, cells and circuits in the peripheral and central nervous system that drive acute and chronic itch transmission. We also discuss novel candidate signaling molecules in neurons, keratinocytes and immune cells and some of the best chronic itch models for assessing the role of such candidates. Finally, we describe some of the latest therapeutic approaches for treating chronic itch conditions in humans. We refer the readers to reviews that comprehensively cover the itch field1,48.

What molecules mediate acute itch transduction?

Itch-sensitive neurons are often classified according to their sensitivity to histamine, the best-studied pruritogen. Histamine is released from immune cells (mast cells and basophilic granulocytes) and keratinocytes and acts on a subset of sensory neurons that express histamine receptors and the heat- and capsaicin-gated ion channel transient receptor potential cation channel, subfamily V, member 1 (TRPV1; Fig. 1). A large body of evidence demonstrates that TRPV1 is a key transduction channel that functions downstream of histamine receptors to induce neuronal excitation and promote itch behaviors; first, TRPV1 antagonists inhibit histamine-evoked signals in DRG neurons; second, patients with allergic rhinitis display increased itch sensations following TRPV1 stimulation; and third, mice lacking TRPV1-expressing neurons or TRPV1 itself display substantially less histamine-induced scratching than wild-type littermates9,10. However, histamine-evoked itch is not completely attenuated in TRPV1-deficient mice10, suggesting that other channel(s) may be required for histaminergic itch11. Interestingly, not all histamine-sensitive neurons are capsaicin-sensitive, but whether this subset of histaminergic neurons mediates itch, pain and/or inflammation, is unknown.

Many forms of itch are insensitive to antihistamine treatment. As such, much research over the last five years has focused on elucidating the mechanisms of nonhistaminergic itch. The exogenous pruritogens cowhage, chloroquine and imiquimod and the endogenous pruritogens serotonin, endothelin-1, thymic stromal lymphopoietic protein (TSLP) and bovine adrenal medulla peptide (BAM) 8–22 signal via distinct histamine-independent molecular pathways and evoke robust scratching behaviors in TRPV1-deficient mice4,6,10,1214 (Table 1). However, the pharmacologic or genetic silencing of TRPV1-expressing neurons attenuates the itch behaviors induced by all of these pruritogens10,12,13. These data suggest that TRPV1-positive afferents mediate itch in response to diverse pruritogens and express a wide variety of itch receptors, signaling molecules and transduction channels.

Table 1

Receptors and channels that mediate itch transduction in sensory neurons

ReceptorAgonistsType
Peripheral Histamine receptor 1 (ref. 68)HistamineGPCR
MrgprA3 (ref. 15)ChloroquineGPCR
MrgprCll (MrgprX1 in humans)15BAM, SLIGRLGPCR
MrgprD16β-AlanineGPCR
Endothelin-1 (refs. 69-71)EndothelinGPCR
5-HT receptors 1-7 (ref. 72)Serotonin, 12(S)-HPETEGPCR
Protease-activated receptor 2 (refs. 24,26,73-75)Tryptase, trypsin, SLIGRL, SLIGR, LIGRLOGPCR
TRL3 (ref. 29)Poly(I:C)Cytokine
TLR7 (ref. 28)ImiquimodCytokine
TSLPR (in complex with IL-7Ra)12TSLPCytokine
IL-31R66IL-31Cytokine
TRPA1 (ref. 14)Allyl isothiocyanate, cinnamaldehyde (reviewed in ref. 76)Ion channel
TRPV1 (refs. 9,10)Capsaicin (reviewed in ref. 77)Ion channel
Central GRPR46,47GRPGPCR
Npra48BMP (NPPB)Receptor GC
MOR1D57OpioidsGPCR

The Mas-related G protein–coupled receptor (Mrgpr) family has emerged as a novel class of histamine-independent itch receptors6,15. MrgprA3 is a receptor for the antimalarial drug chloroquine, which has a side effect of itch15; but the endogenous agonist(s) that activate MrgprA3 in vivo to induce itch is not known. MrgprC11 is a receptor for the mast cell amide BAM. MrgprD is a receptor for β-alanine, which induces itch, mild prickling, tingling and burning sensations16. The MrgprD neurons are distinct from other classes of itch neurons in that they do not all express TRPV1 or TRPA1 ion channels and have been linked to pain, suggesting a more complex role for these neurons in somatosensation1721. Chloroquine- and BAM-evoked itch also requires the irritant receptor TRPA1, which acts as the primary transduction channel downstream of MrgprA3 and MrgprC11. TRPA1 is required for both chloroquine- and BAM-evoked depolarization of somatosensory neurons as well as chloroquine- and BAM-evoked scratching in mice14. Interestingly, although signaling downstream of these two itch receptors converges on TRPA1, MrgprC11 signals through phospholipase C (PLC) and MrgprA3 via Gβϒ (ref. 14). These studies highlight a potential role for Mrgprs and TRPA1 channels as mediators of chronic itch.

Patients with a variety of chronic itch conditions display high levels of proteases in their skin, which can activate protease receptors, including protease activated receptor 2 (PAR2)22. PAR2 is expressed in both keratinocytes and DRG neurons and is proposed to promote acute and chronic itch2224. A direct role for PAR2 in neuronal itch signaling was suggested by studies using the ligand Ser-Leu-Ile-Gly-Arg-Leu (SLIGRL), which was proposed to act as a mimetic of the PAR2 ligand that is generated by proteolytic cleavage of the tethered peptide in the amino-terminal extracellular domain of PAR2 (ref. 24). When injected intradermally, SLIGRL induces neuronal excitability and robust itch behaviors in rodents. However, a recent study showed that SLIGRL also activates the MrgprC11 receptor, and scratch responses triggered by this peptide are abolished in Mrgpr-null animals but not affected in PAR2-deficient mice25. However, there is strong evidence for a role of PAR2 in human and mouse models of chronic itch. For example, tryptase and cathepsin S activate PAR2 by cleaving the N terminus, and their proteolytic products are increased in the skin of patients with atopic dermatitis23,26. Interestingly, the endogenous generation of PAR2 agonist by tryptase triggers robust itch behaviors that are attenuated in PAR2-deficient mice12. Unlike SLIGRL, tryptase does not activate rodent MrgprC11 or human MrgprX1, suggesting that SLIGRL is not equivalent to the natural tethered ligand in vivo12. In fact, PAR2 activation by SLIGRL, tryptase or cathepsin S also promotes the release of the cytokine, thymic stromal lymphopoietin (TSLP) from keratinocytes, indicating that intricate multicellular pathways may underlie chronic itch12,27. TSLP is a robust pruritogen associated with atopic dermatitis that activates a subset of neurons expressing TRPA1- and TSLP receptor (TSLPR) (Fig. 1). Indeed, TRPA1-deficient mice display significantly reduced itch behaviors after TSLP injection12. The broad expression of PAR2 and TSLPR in neurons, epithelial cells and immune cells, suggests that the analysis of tissue-specific knockouts is needed to elucidate the relative contributions of these cell types to itch.

Recent studies have implicated Toll-like receptors (TLRs) in neuronal itch signaling2830. TLRs are best known for their role in innate immune responses. These membrane proteins act as sentinels for components of foreign organisms such as lipopolysaccharides, lipo-proteins, flagellin, double- and single-stranded RNA and CpG DNA, and they also bind endogenous compounds released by necrosis or tissue remodeling31. TLRs are broadly expressed and are associated with pathological inflammatory conditions32,33. A number of reports suggest that TLRs have a key role in itch. First, tissue profiling suggests that sensory neurons express TLR3, TLR4, TLR7 and TLR9, but different studies describe distinct expression profiles34. Second, TLR agonists promote activation of sensory neurons, but whether they act directly on TLRs has not been settled, making the exact mechanism by which these receptors function in itch unclear13,28,35. However, TLRs remain intriguing candidates, as there are links between TLR polymorphisms and susceptibility to chronic itch disorders36,37. Owing to the broad expression of TLRs in keratinocytes, neurons and immune cells, tissue-specific knockout of candidate genes is required to establish the exact roles of TLRs in chronic itch.

Are itch neurons a homogeneous or heterogeneous population?

The number of itch-neuron subtypes remains enigmatic. Studies using the plant-derived pruritogen cowhage were the first to propose that there are histamine-dependent and histamine-independent itch pathways in the periphery4. Consistent with this model, many forms of chronic itch are insensitive to antihistamine treatment. Recent ablation studies also support a model with at least two subtypes of itch neuron. Genetic ablation of MrgprA3-expressing neurons attenuates mouse itch behaviors evoked by many pruritic compounds; responses to chloroquine, a direct MrgprA3 agonist, were nearly abolished, and itch behaviors to other pruritogens were partially reduced38. Likewise, genetic ablation of calcitonin gene–related peptide (CGRP)-expressing sensory neurons (which includes the MrgprA3 subgroup) resulted in a near complete loss of chloroquine- and histamine-evoked itch behaviors39. These results suggest that MrgprA3-positive neurons have an important function in itch but that additional neuronal subsets also detect pruritic stimuli (Fig. 1). Finally, pharmacological silencing of distinct subsets of neurons also suggests that itch-sensitive neurons are heterogeneous, with partially overlapping populations of histamine- and chloroquine-responsive cells40.

The existence of separate itch and pain pathways is also controversial. Indeed, there has been a longstanding debate over how input from somatosensory neurons gives rise to the distinct sensations of itch and pain, leading to a number of models. First, the intensity model suggests that itch neurons are polymodal and mediate sensations of both itch and pain; itch sensations are triggered when these neurons are weakly activated, whereas pain sensations are elicited when neurons are strongly activated. Second, the spatial-contrast model suggests that itch and pain neurons are polymodal but can be distinguished by their differential localization in the skin. Third, the labeled-line model proposes that itch and pain pathways are functionally and anatomically segregated in the PNS and CNS. Finally, the selectivity model posits that itch and pain neurons are polymodal, with itch sensations occurring when selective subsets of itch neurons are activated alone and pain sensations dominating when itch and pain neurons are activated together.

A number of studies support models in which itch neurons are polymodal. All itch-sensitive neurons also respond to classical algogenic stimuli, such as heat and capsaicin6. Likewise, high doses of pruritogens, such as histamine, induce pain, and low doses of algogens, such as capsaicin, evoke itch6. In further support of this model, electrophysiological recordings show that many peripheral and central neurons respond to both itch and pain stimuli4143. The selectivity theory is supported by evidence that blocking pain transmission through deletion of the gene encoding vesicular glutamate transporter 2 (VGLUT2) in a subset of nociceptors triggers spontaneous itch behaviors44,45. However, whether the acute block of pain transmission also unmasks itch circuit activity is unknown.

A number of studies also support the existence of a selective subpopulation of itch neurons (Table 2). First, spinal neurons that express gastrin-releasing peptide receptor (GRPR) and the natriuretic polypeptide receptor subtype A (NPRA) are required for histamine-dependent and histamine-independent itch behaviors but not pain behaviors4648. Second, a population of dorsal horn inhibitory interneurons expressing the basic helix-loop-helix B5 (Bhlhb5) transcription factor suppress itch but not pain-signal transmission5. Third, mice lacking MrgprA3-positive neurons display attenuated itch behaviors but normal pain sensitivity38. Fourth, mice expressing TRPV1 solely in MrgprA3-positive neurons exhibit scratching, but no pain behaviors when challenged with the TRPV1 agonist capsaicin38. Lastly, the pharmacological silencing of itch- sensitive neurons has no effect on pain behavior40. Overall, these experiments suggest that detection of itch and pain stimuli at the periphery occur through distinct classes of neurons and that the itch circuit is, at least in part, separate from the pain circuit.

Table 2

The cellular specificity of itch

Loss of behavioral responses
Response to somatosensory stimuli
Subset of sensory neurons ablated or silencedHistamineChloroquinePAR2β-AlanineHeatPainTouch
Capsaicin-induced ablation of all TRPV1 neurons10,18Complete lossComplete lossComplete lossNDComplete lossComplete lossNormal
Molecular genetic ablation of all CGRP-expressing neurons: ~50% of TRPV1 neurons eliminated39Complete lossComplete lossNDNDComplete lossComplete lossNormal
Molecular genetic ablation of all MrgA3-expressing neurons: ~20% TRPV1 neurons eliminated38Partial lossComplete lossPartial lossNDNormalNormalNormal
QX314-evoked silencing of chloroquine/PAR2-sensitive neurons40NormalComplete lossComplete lossNDNormalNormalNormal
QX314-evoked silencing of histamine-sensitive neurons40Complete lossNormalNormalNDNDNormalNormal

These studies demonstrate that particular subsets of neurons are required for the detection of the itch-inducing compounds histamine, chloroquine and PAR2 and other somatosensory behaviors, including touch, heat and pain. Complete loss, complete loss of a somatosensory behavior; partial loss, partial elimination; normal, no change; ND, not determined (modality was not tested).

Among these studies, there are some discrepancies in the percentages of cells required for itch responses as well as in the exact molecular identity of the cells responsible for itch (Table 2). Such differences may be due to varying experimental conditions or age of animals, incomplete ablation of neuronal subtypes or the timing of ablation or silencing relative to experimentation. Likewise, a number of unanswered questions about peripheral itch transduction persist, including the mechanisms by which itch is induced by the pruritogens serotonin, endothelin-1 and imiquimod (Table 1). Future studies will more precisely define the molecules and diverse populations of neurons required for itch detection.

How are itch signals transmitted to the spinal cord?

A number of exciting advances have been made in characterizing and identifying itch neurotransmitters. Glutamate was originally thought to be the main neurotransmitter released by primary itch neurons. Recent studies have suggested that gastrin releasing peptide (GRP), B-type natriuretic peptide (BNP; also called naturietic polypeptide B (NPPB)) and neuromedin B (NMB) may act as itch transmitters in conjunction with glutamate4649. Whether NMB is a selective neurotransmitter is unknown; one study showed a role for NMB in itch50, whereas another showed that NMB mediates pain but not itch transmission49. However, GRP and BNP are thought to function as itch-specific transmitters.

Evidence for glutamate acting as co-transmitter came from studies in which glutamate transmission was abolished in sensory neurons through deletion of the gene encodingVGLUT2 in TRPV1-expressing neurons44,45. Mice lacking VGLUT2 showed enhanced itch responses, suggesting that an additional neurotransmitter(s) is used at the first itch synapse in the absence of glutamate, and that glutamate-mediated transmission is required for disinhibition of itch (discussed below).

GRP has emerged as a key mediator of itch transmission. Ablation of dorsal horn interneurons that express GRPR attenuates itch behaviors in response to a variety of pruritogens, with no change in pain behaviors47. This seminal study suggested that itch and pain signals are transmitted via distinct circuits. However, whether GRP acts as a peripheral or central neurotransmitter remains controversial. Several lines of evidence support a model wherein GRP acts as a peripheral itch neurotransmitter that signals to second-order GRPR-expressing neurons in the spinal cord (Fig. 2, model 1). First, a number of studies have shown GRP expression in a subset of primary afferent sensory neurons and GRPR staining in dorsal spinal neurons46,47,51. Second, MrgprA3-positive primary afferent nerve terminals colocalize with synapsin and GRPR-expressing cells in the dorsal horn of the spinal cord38. Third, GRP expression was significantly increased in primary afferents in mice that display enhanced spontaneous and evoked itch behaviors due to constitutive activity of BRAF kinase in nociceptors52. However, a number of studies have reported little or no GRP staining in primary afferents48,49,51 but robust GRP expression in dorsal horn neurons48,51,53. Indeed, GRP was first described in a microarray screen to identify dorsal horn–enriched genes54. Despite the contradictory evidence of GRP expression in peripheral tissue and the unresolved role of GRP in spinal cord neurons, GRP clearly has a key function in itch transmission, as mice deficient in GRPR do not scratch in response to a variety of pruritogens; however, these mice display normal itch responses to histamine, endothelin-1 and serotonin, suggesting the existence of multiple itch transmitters46.

An external file that holds a picture, illustration, etc.
Object name is nihms-671116-f0002.jpg

Models of peripheral and spinal itch circuitry. (a) Excitatory itch pathways. Sensory neurons have cell bodies in the DRG (blue) and project primary afferents to the skin, where a variety of transduction molecules reside and detect itch, touch and pain stimuli. These neurons also send projections to the dorsal horn of the spinal cord, where they synapse with secondary neurons (gray). Second-order spinal neurons connect with other dorsal horn interneurons and/or spinal thalamic projection neurons (green). Model 1 proposes that GRP is a peripheral itch transmitter that activates GRPR on secondary pruriceptors. Chloroquine, SLIGRL and the mast-cell degranulating compound 48/80 (48/80) all require functional GRP signaling to evoke itch. Histamine (Hist), 5-hydroxytryptamine (5-HT) and endothelin-1 induce itch independent of GRP signaling. Model 2 proposes that BNP serves as a peripheral itch transmitter that activates secondary pruriceptors that express NPRA and GRP. These neurons then activate tertiary GRPR-expressing pruriceptors. Model 3 is a combination of models 1 and 2, in which BNP and GRP act in parallel as peripheral itch transmitters. GRP can be released by DRG neurons and NPRA-positive spinal neurons. BNP activates NPRA-expressing spinal interneurons, and GRP activates GRPR-expressing spinal interneurons. (b) Model for the inhibitory inputs of the itch pathway whereby tonic activity is inhibited by paired activation of the pain circuitry. A subset of VGLUT2-expressing nociceptors in the DRG and a subset of dorsal horn inhibitory interneurons expressing Bhlhb5 specifically suppress itch signal transmission. Connections that are unknown or controversial are shown as dashed magenta arrows. CQ, chloroquine.

Recent studies have implicated BNP, a neurohormone secreted by the cardiac ventricles in response to volume expansion and pressure overload55, in itch and pain transmission48,56. BNP is the modified product of the gene Nppb and targets the receptor NPRA. BNP is expressed in a subset of primary afferent neurons, and intrathecal injection of BNP triggers robust itch in mice and attenuation of inflammatory pain in rats. Mice lacking BNP display a loss of itch behaviors to a wide variety of pruritogens. Likewise, ablation of NPRA-expressing interneurons eliminated BNP- and histamine-evoked itch behaviors. Overall these experiments demonstrate a key role for BNP in itch transmission.

What is the relationship between the GRP and BNP signaling pathways? Two findings support a model whereby BNP release by primary afferent neurons triggers the release of GRP by spinal cord interneurons (Fig. 2, model 2). First, NPRA is expressed in a subset of spinal interneurons that also express GRP. Second, mice lacking NPRA-expressing interneurons display normal GRP-evoked itch. Nevertheless, these experiments do not exclude the possibility that GRP may act in parallel with BNP to transmit signals from the PNS to the CNS (Fig. 2, model 3). Distinguishing between these three models will be an essential next step in understanding itch signal transmission.

How are itch signals modulated by spinal interneurons?

Some noxious mechanical (for example, scratching) and thermal stimuli (cold) inhibit the sensation of itch, whereas light stroking potentiates itch. Indeed, recent studies have shed light on our understanding of how spinal inhibitory interneurons contribute to itch sensations and how spinal thalamic tract (STT) neurons integrate multiple signals.

Excitatory pathways

Recent studies have identified three excitatory itch-interneuron subsets that express morphine receptors, the transcription factor T cell leukemia homeobox 3 (Tlx3) or the testicular orphan nuclear receptor (TR4). The analgesic morphine produces a side effect of intense itch, which can be reversed with μ-opioid receptor antagonists57. Several studies suggest that morphine-evoked itch and analgesia are mediated by two independent mechanisms. First, genetic analyses revealed that itch responses are dependent on MOR1D, a splice variant of the μ-opioid receptor, whereas analgesia requires the canonical form, MOR1 (ref. 57). The MOR1D variant is coexpressed with GRPR, suggesting that these receptors may form heterodimers that mediate morphine-evoked itch in a subset of morphine-sensitive interneurons. Second, electrophysiological analyses of trigeminothalamic tract neurons revealed that morphine increases the activity of pruriceptive interneurons and decreases activity of nociceptive interneurons58.

The deletion of TR4 eliminates a subset of superficial dorsal horn neurons, which leads to a profound unresponsiveness to pruritogens53. Neurons that express both GRP and NPRA and neurons that express GRPR are missing in TR4-deficient mice. In contrast, the number of inhibitory dorsal horn neurons, primary afferent neurons and projection neurons are normal in TR4- deficient mice. Likewise, mice that lack Tlx3 in dorsal horn excita-tory neurons have major pruriceptive deficits owing to a loss of excitatory neurons that express somatostatin, preprotachykinin 1, GRP and/or GRPR59. Overall, these studies show that distinct subsets of dorsal horn neurons are required for itch transmission. However, these experiments do not address the issue of itch specificity in the CNS, as TR4- and Tlx3-deficient mice also display defects in pain behaviors.

Inhibitory pathways

Genetic ablation studies suggest that tonic activity of the itch circuit is inhibited by activation of the pain circuit. The genetic deletion of VGLUT2 in nociceptors abolishes pain transmission and leads to robust spontaneous itch behaviors44,45. Similarly, mice lacking a subset of inhibitory dorsal spinal neurons owing to the conditional deletion of the Bhlhb5 transcription factor5 also display spontaneous itch behaviors. In addition to spontaneous scratching, both VGLUT2- and Bhlhb5-deficient mice display augmented pruritic responses to itch-inducing agents5,44,45. Together, these studies support a model of cross-inhibition, whereby glutamatergic pain afferents activate Bhlhb5-expressing interneurons, which in turn inhibit the itch circuitry (Fig. 2). This cross-inhibition model is also supported by the observation that activity-dependent silencing of neurons expressing TRPA1 or TRPV1 switches the behavioral response to mustard oil or capsaicin application from pain to itch40. The roles of known neurotransmitters60 and other signaling mechanisms in cross-inhibition remain unknown. However, the κ-opioid receptor agonist nalfurafine inhibits scratching, suggesting that κ-opioid receptor–expressing neurons may form part of the inhibitory pathway for itch61.

How are itch signals sent to and processed by the brain?

Electrophysiological itch studies of spinal thalamic tract (STT) neurons have examined the effects of cowhage and histamine on excitability. These studies show that cowhage and histamine elicit responses in separate populations of neurons, suggesting that different pruritogens activate distinct central pathways4,43,62. In addition, these studies show there is a convergence of itch and pain signals before the STT neurons. For example, itch-evoked firing of STT neurons in primates is enhanced by stroking, and inhibited by scratching, the skin40.

Like STT recordings, brain imaging shows that cowhage and histamine activate distinct brains regions as well as some overlapping areas63. How this anatomical segregation affects the perception of itch is unclear. Future studies will define the brain regions that process itch sensations and may help explain the paradoxical convergence of itch and pain signals and apparent segregation of different pruriceptive stimuli.

What molecules and cells contribute to chronic itch?

Studies on acute itch have led to the identification of candidate molecules and cell types that may represent good targets for treating chronic itch. A number of chronic itch models have recently been developed to probe the function of transgenic mice lacking such candidates. The different models mimic distinct human itch disorders and fall into three broad categories: chemically induced itch, genetically induced itch or chronic itch that develops naturally in some inbred mouse strains (Table 3).

Table 3

Chronic itch models in rodents that recapitulate human chronic itch disorders

Itch assay or mouse lineHuman disease modelSpeciesGenetic modification
Chemical AEW78,79Dry skinMouseTrpa1−/−, Mrgpra3 ablation
Ovalbumin sensitization80AllergyMouseTrpa1−/−, Mrgpra3 ablation
Oxazolone80Irritant contact dermatitisMouseNone
Ethynylestradiol60CholestasisRatNone
Diphenylcyclopropenone81Contact dermatitisMouseNone
2,4,6-Trinitro-1-chlorobenzene80Contact dermatitisMouseNone
Dorsal horn quisqualate injection82Neuropathic itchRatNone
Streptozotocin83DiabetesRatNone
Dermatophagoides pteronyssinus 84 Allergy; asthmaMouseNone
Genetic K5 IL-13 (IL-13 overexpression in skin)85Atopic dermatitisMouseNone
K14 human chymotryptase overexression86PsoriasisMouseNone
K5 IL-4 overexpression80Atopic dermatitisMouseNone
Cathepsin S overexpression80Atopic dermatitisMouse Par2−/−
K5IL-18Tg (IL-18 overexpression in skin)87Atopic dermatitisMouseNone
KCASP1Tg (caspase 1 overexpression in skin)87Atopic dermatitisMouseNone
Spink5−/− (ref. 88)Netherton syndromeMouseNone
TLR4-deficient mice and 3-week Aspergillus fumigatus allergen89Atopic dermatitisMouseNone
Stat6VT and vitamin D90Atopic dermatitisMouseNone
SPT-cKO mice (knockout of serine palmitoyltransferase, critical for ceramide biosynthesis)91PsoriasisMouseNone
IL-31 overexpression80Atopic dermatitisMouseNone
TSLP overexpression80,92,93Atopic dermatitisMouseNone
IL-18 overexpression80Atopic dermatitisMouseNone
RelB−/− (ref. 80)Atopic dermatitisMouseNone
Ctse−/− (ref. 80)Atopic dermatitisMouseNone
Msx2-cre Rbpj−/− (DNA-binding partner of Notch)93Barrier disruptionMouseNone
K5-TSLP80Atopic dermatitisMouseNone
K5-tTA-Tight-IL-13 (IL-13 overexpression in skin)94Atopic dermatitisMouse Trpa1−/−
Inbred strain DAbg/bg80Spontaneous dermatitis; atopic dermatitisRatNone
NC/Tnd mice/NGA80Atopic dermatitisMouseNone
SJL/J mice95Chronic inflammatory liver diseaseMouseNone
Flaky tail mice80Atopic dermatitisMouseNone
Naruto Research Institute Otsuka Atrichia (NOA) mice80Atopic dermatitisMouseNone
Other Bile duct ligation96Liver diseaseRatNone
Scratching97Scratching; itch-scratch cycleMouseNone
Wire-brush scratching98Barrier disruptionMouseNone

The discovery of MrgprA3, MrgprC11 and TRPA1 as histamine-independent itch transducers led to a number of studies examining the roles of these receptors in chronic itch. The importance of MrgprA3-positive neurons was examined in both the acetone ether water (AEW) dry skin–evoked model of itch and the ovalbumin-induced allergic contact dermatitis model38. In the AEW model, mice deficient for these neurons exhibited a dramatic decrease in itch behaviors, to levels similar to that observed in untreated mice38. Loss of MrgprA3-positive neurons also led to a 50% decrease in allergic itch behaviors, suggesting that these neurons contribute at least partially to allergic itch38. The different outcomes of these chronic itch models is not surprising, as distinct forms of chronic itch in humans respond to different treatments and thus are hypothesized to involve distinct mediators. Mice lacking TRPA1 also displayed significantly decreased itch behaviors in both the AEW and allergic itch models34. In addition to decreased itch behaviors, loss of TRPA1 attenuated epidermal thickening and changes in gene expression in both neurons and skin that normally accompany the itch-scratch cycle34. In the allergic model, loss of TRPA1 decreased skin edema, immune cell infiltration and inflammatory cytokine levels64. These studies highlight a role for TRPA1 in both itch signal transmission and in the pathophysiological changes that occur within the skin to promote and/or maintain chronic itch.

Many studies using chronic itch models have focused on the communication between skin cells and the immune cells that promote itch and inflammation. Indeed, the use of immune cell inhibitors that target mast cells, basophils or T lymphocytes have been effective at treating antihistamine-insensitive chronic itch conditions. For example, in a clinical trial, monoclonal anti-IgE antibodies provided relief from the debilitating itch caused by idiopathic urticaria65. Likewise, a recent study using anti IL-31 antibodies was successful in treating inflammatory responses in atopic dermatitis66. Importantly, these trials highlight a key role for the immune system in mediating chronic itch. In addition to immune cells, sensory neuron terminals and epithelial cells also release inflammatory mediators that contribute to a complex network of multi-directional communication between cell types. For example, sensory neurons release substance P, which promotes immune cell chemoattraction and activation67. Similarly, sensory neurons and epithelial cells express receptors for inflammatory mediators28,29,34. Future studies using diverse itch models and tissue-specific knockout mice will help define the roles of each of these cell types in chronic itch.

Where do we go from here?

A number of key questions remain to be addressed to understand the development and maintenance of chronic itch. First, whether the molecules and cells that mediate acute itch, chronic itch and pain are the same remains controversial. Second, the changes in gene expression and somatosensory circuitry that occur during the development and maintenance of chronic itch have yet to be discovered. Third, it is unknown whether modulation of dorsal horn interneurons in the itch circuit promotes neuropathic itch. Fourth, the signaling mechanisms and relative contributions of epithelial cells, immune cells and sensory neurons in chronic itch must be elucidated. Finally, candidate mechanisms must be validated in a variety of different human itch disorders and translated to the clinic so that effective therapeutic strategies can alleviate a variety of human itch disorders. Identifying the molecules and cells that are shared and those that are unique to the itch and pain pathways may lead to the development of drugs and therapies to treat itch without affecting pain and to alleviate pain without triggering itch. The dramatic progress in recent molecular and clinical studies suggests better therapies for chronic itch are close at hand.

ACKNOWLEDGMENTS

We are grateful to M. Pellegrino for critical comments on the manuscript. This work was supported by US National Institutes of Health (NIH) grants AR059385 and DOD007123A (D.M.B.), the intramural research program of the NIH NIDCR (M.A.H.) and a US National Science Foundation Fellowship (S.R.W.).

Footnotes

COMPETING FINANCIAL INTERESTS

The authors declare competing financial interests: details are available at http://www.nature.com/doifinder/10.1038/nn.3619.

References

1. Ikoma A, Steinhoff M, Stander S, Yosipovitch G, Schmelz M. The neurobiology of itch. Nat. Rev. Neurosci. 2006;7:535–547. [PubMed] [Google Scholar]
2. Patel T, Yosipovitch G. Therapy of pruritus. Expert Opin. Pharmacother. 2010;11:1673–1682. [PMC free article] [PubMed] [Google Scholar]
3. Elmariah SB, Lerner EA. Topical therapies for pruritus. Semin. Cutan. Med. Surg. 2011;30:118–126. [PMC free article] [PubMed] [Google Scholar]
4. Davidson S, Giesler GJ. The multiple pathways for itch and their interactions with pain. Trends Neurosci. 2010;33:550–558. [PMC free article] [PubMed] [Google Scholar]
5. Ross SE. Pain and itch: insights into the neural circuits of aversive somatosensation in health and disease. Curr. Opin. Neurobiol. 2011;21:880–887. [PubMed] [Google Scholar]
6. McNeil B, Dong X. Peripheral mechanisms of itch. Neurosci. Bull. 2012;28:100–110. [PMC free article] [PubMed] [Google Scholar]
7. Akiyama T, Carstens E. Neural processing of itch. Neuroscience. 2013;250:697–714. [PMC free article] [PubMed] [Google Scholar]
8. Garibyan L, Rheingold CG, Lerner EA. Understanding the pathophysiology of itch. Dermatol. Ther. 2013;26:84–91. [PMC free article] [PubMed] [Google Scholar]
9. Shim WS, et al. TRPV1 mediates histamine-induced itching via the activation of phospholipase A2 and 12-lipoxygenase. J. Neurosci. 2007;27:2331–2337. [PMC free article] [PubMed] [Google Scholar]
10. Imamachi N, et al. TRPV1-expressing primary afferents generate behavioral responses to pruritogens via multiple mechanisms. Proc. Natl. Acad. Sci. USA. 2009;106:11330–11335. [PMC free article] [PubMed] [Google Scholar]
11. Akiyama T, Carstens MI, Carstens E. Enhanced scratching evoked by PAR-2 agonist and 5-HT but not histamine in a mouse model of chronic dry skin itch. Pain. 2010;151:378–383. [PMC free article] [PubMed] [Google Scholar]
12. Wilson SR, et al. The epithelial cell-derived atopic dermatitis cytokine TSLP activates neurons to induce itch. Cell. 2013;155:285–295. [PMC free article] [PubMed] [Google Scholar]
13. Kim SJ, et al. Analysis of cellular and behavioral responses to imiquimod reveals a unique itch pathway in transient receptor potential vanilloid 1 (TRPV1)-expressing neurons. Proc. Natl. Acad. Sci. USA. 2011;108:3371–3376. [PMC free article] [PubMed] [Google Scholar]
14. Wilson SR, et al. TRPA1 is required for histamine-independent, Mas-related G protein–coupled receptor-mediated itch. Nat. Neurosci. 2011;14:595–602. [PMC free article] [PubMed] [Google Scholar]
15. Liu Q, et al. Sensory neuron-specific GPCR Mrgprs are itch receptors mediating chloroquine-induced pruritus. Cell. 2009;139:1353–1365. [PMC free article] [PubMed] [Google Scholar]
16. Liu Q, et al. Mechanisms of itch evoked by β-alanine. J. Neurosci. 2012;32:14532–14537. [PMC free article] [PubMed] [Google Scholar]
17. Zylka MJ, Rice FL, Anderson DJ. Topographically distinct epidermal nociceptive circuits revealed by axonal tracers targeted to Mrgprd. Neuron. 2005;45:17–25. [PubMed] [Google Scholar]
18. Cavanaugh DJ, et al. Distinct subsets of unmyelinated primary sensory fibers mediate behavioral responses to noxious thermal and mechanical stimuli. Proc. Natl. Acad. Sci. USA. 2009;106:9075–9080. [PMC free article] [PubMed] [Google Scholar]
19. Rau KK, et al. Mrgprd enhances excitability in specific populations of cutaneous murine polymodal nociceptors. J. Neurosci. 2009;29:8612–8619. [PMC free article] [PubMed] [Google Scholar]
20. Shields SD, Cavanaugh DJ, Lee H, Anderson DJ, Basbaum AI. Pain behavior in the formalin test persists after ablation of the great majority of C-fiber nociceptors. Pain. 2010;151:422–429. [PMC free article] [PubMed] [Google Scholar]
21. Mishra SK, Tisel SM, Orestes P, Bhangoo SK, Hoon MA. TRPV1-lineage neurons are required for thermal sensation. EMBO J. 2011;30:582–593. [PMC free article] [PubMed] [Google Scholar]
22. Steinhoff M, et al. Proteinase-activated receptor-2 in human skin: tissue distribution and activation of keratinocytes by mast cell tryptase. Exp. Dermatol. 1999;8:282–294. [PubMed] [Google Scholar]
23. Steinhoff M, et al. Proteinase-activated receptor-2 mediates itch: a novel pathway for pruritus in human skin. J. Neurosci. 2003;23:6176–6180. [PMC free article] [PubMed] [Google Scholar]
24. Shimada SG, Shimada KA, Collins JG. Scratching behavior in mice induced by the proteinase-activated receptor-2 agonist, SLIGRL-NH2. Eur. J. Pharmacol. 2006;530:281–283. [PubMed] [Google Scholar]
25. Liu Q, et al. The distinct roles of two GPCRs, MrgprC11 and PAR2, in itch and hyperalgesia. Sci. Signal. 2011;4:ra45. [PMC free article] [PubMed] [Google Scholar]
26. Reddy VB, Shimada SG, Sikand P, Lamotte RH, Lerner EA. Cathepsin S elicits itch and signals via protease-activated receptors. J. Invest. Dermatol. 2010;130:1468–1470. [PMC free article] [PubMed] [Google Scholar]
27. Kim N, et al. Overexpression of cathepsin S induces chronic atopic dermatitis in mice. J. Invest. Dermatol. 2012;132:1169–1176. [PubMed] [Google Scholar]
28. Liu T, Xu ZZ, Park CK, Berta T, Ji RR. Toll-like receptor 7 mediates pruritus. Nat. Neurosci. 2010;13:1460–1462. [PMC free article] [PubMed] [Google Scholar]
29. Liu T, et al. TLR3 deficiency impairs spinal cord synaptic transmission, central sensitization, and pruritus in mice. J. Clin. Invest. 2012;122:2195–2207. [PMC free article] [PubMed] [Google Scholar]
30. Liu T, Gao YJ, Ji RR. Emerging role of Toll-like receptors in the control of pain and itch. Neurosci. Bull. 2012;28:131–144. [PMC free article] [PubMed] [Google Scholar]
31. Barton GM, Kagan JC. A cell biological view of Toll-like receptor function: regulation through compartmentalization. Nat. Rev. Immunol. 2009;9:535–542. [PMC free article] [PubMed] [Google Scholar]
32. Abreu MT, Fukata M, Arditi M. TLR signaling in the gut in health and disease. J. Immunol. 2005;174:4453–4460. [PubMed] [Google Scholar]
33. Sabroe I, Parker LC, Wilson AG, Whyte MK, Dower SK. Toll-like receptors: their role in allergy and nonallergic inflammatory disease. Clin. Exp. Allergy. 2002;32:984–989. [PubMed] [Google Scholar]
34. Wilson SR, et al. The ion channel TRPA1 is required for chronic itch. J. Neurosci. 2013;33:9283–9294. [PMC free article] [PubMed] [Google Scholar]
35. Chiu IM, et al. Bacteria activate sensory neurons that modulate pain and inflammation. Nature. 2013;501:52–57. [PMC free article] [PubMed] [Google Scholar]
36. Salpietro C, et al. TLR2 and TLR4 gene polymorphisms and atopic dermatitis in Italian children: a multicenter study. Int. J. Immunopathol. Pharmacol. 2011;24:33–40. [PubMed] [Google Scholar]
37. Mrabet-Dahbi S, et al. The Toll-like receptor 2 R753Q mutation modifies cytokine production and Toll-like receptor expression in atopic dermatitis. J. Allergy Clin. Immunol. 2008;121:1013–1019. [PubMed] [Google Scholar]
38. Han L, et al. A subpopulation of nociceptors specifically linked to itch. Nat. Neurosci. 2013;16:174–182. [PMC free article] [PubMed] [Google Scholar]
39. McCoy ES, et al. Peptidergic CGRPα primary sensory neurons encode heat and itch and tonically suppress sensitivity to cold. Neuron. 2013;78:138–151. [PMC free article] [PubMed] [Google Scholar]
40. Roberson DP, et al. Activity-dependent silencing reveals functionally distinct itch-generating sensory neurons. Nat. Neurosci. 2013;16:910–918. [PMC free article] [PubMed] [Google Scholar]
41. Namer B, et al. Separate peripheral pathways for pruritus in man. J. Neurophysiol. 2008;100:2062–2069. [PMC free article] [PubMed] [Google Scholar]
42. Johanek LM, et al. A role for polymodal C-fiber afferents in nonhistaminergic itch. J. Neurosci. 2008;28:7659–7669. [PMC free article] [PubMed] [Google Scholar]
43. Davidson S, et al. Pruriceptive spinothalamic tract neurons: physiological properties and projection targets in the primate. J. Neurophysiol. 2012;108:1711–1723. [PMC free article] [PubMed] [Google Scholar]
44. Lagerström MC, et al. VGLUT2-dependent sensory neurons in the TRPV1 population regulate pain and itch. Neuron. 2010;68:529–542. [PMC free article] [PubMed] [Google Scholar]
45. Liu Y, et al. VGLUT2-dependent glutamate release from nociceptors is required to sense pain and suppress itch. Neuron. 2010;68:543–556. [PMC free article] [PubMed] [Google Scholar]
46. Sun YG, Chen ZF. A gastrin-releasing peptide receptor mediates the itch sensation in the spinal cord. Nature. 2007;448:700–703. [PubMed] [Google Scholar]
47. Sun YG, et al. Cellular basis of itch sensation. Science. 2009;325:1531–1534. [PMC free article] [PubMed] [Google Scholar]
48. Mishra SK, Hoon MA. The cells and circuitry for itch responses in mice. Science. 2013;340:968–971. [PMC free article] [PubMed] [Google Scholar]
49. Mishra SK, Holzman S, Hoon MA. A nociceptive signaling role for neuromedin B. J. Neurosci. 2012;32:8686–8695. [PMC free article] [PubMed] [Google Scholar]
50. Sukhtankar DD, Ko MC. Physiological function of gastrin-releasing peptide and neuromedin B receptors in regulating itch scratching behavior in the spinal cord of mice. PLoS ONE. 2013;8:e67422. [PMC free article] [PubMed] [Google Scholar]
51. Fleming MS, et al. The majority of dorsal spinal cord gastrin releasing peptide is synthesized locally whereas neuromedin B is highly expressed in pain- and itch-sensing somatosensory neurons. Mol. Pain. 2012;8:52. [PMC free article] [PubMed] [Google Scholar]
52. Zhao ZQ, et al. Chronic itch development in sensory neurons requires BRAF signaling pathways. J. Clin. Invest. 2013;123:4769–4780. [PMC free article] [PubMed] [Google Scholar]
53. Wang X, et al. Excitatory superficial dorsal horn interneurons are functionally heterogeneous and required for the full behavioral expression of pain and itch. Neuron. 2013;78:312–324. [PMC free article] [PubMed] [Google Scholar]
54. Li MZ, et al. Molecular mapping of developing dorsal horn-enriched genes by microarray and dorsal/ventral subtractive screening. Dev. Biol. 2006;292:555–564. [PubMed] [Google Scholar]
55. Forssmann WG, et al. The heart is the center of a new endocrine, paracrine, and neuroendocrine system. Arch. Histol. Cytol. 1989;52(suppl.):293–315. [PubMed] [Google Scholar]
56. Zhang FX, et al. Inhibition of inflammatory pain by activating B-type natriuretic peptide signal pathway in nociceptive sensory neurons. J. Neurosci. 2010;30:10927–10938. [PMC free article] [PubMed] [Google Scholar]
57. Liu XY, et al. Unidirectional cross-activation of GRPR by MOR1D uncouples itch and analgesia induced by opioids. Cell. 2011;147:447–458. [PMC free article] [PubMed] [Google Scholar]
58. Moser HR, Giesler GJ., Jr. Itch and analgesia resulting from intrathecal application of morphine: contrasting effects on different populations of trigeminothalamic tract neurons. J. Neurosci. 2013;33:6093–6101. [PMC free article] [PubMed] [Google Scholar]
59. Xu Y, et al. Ontogeny of excitatory spinal neurons processing distinct somatic sensory modalities. J. Neurosci. 2013;33:14738–14748. [PMC free article] [PubMed] [Google Scholar]
60. Akiyama T, Iodi Carstens M, Carstens E. Transmitters and pathways mediating inhibition of spinal itch-signaling neurons by scratching and other counterstimuli. PLoS ONE. 2011;6:e22665. [PMC free article] [PubMed] [Google Scholar]
61. Inan S, Cowan A. Nalfurafine, a kappa opioid receptor agonist, inhibits scratching behavior secondary to cholestasis induced by chronic ethynylestradiol injections in rats. Pharmacol. Biochem. Behav. 2006;85:39–43. [PubMed] [Google Scholar]
62. Davidson S, Zhang X, Khasabov SG, Simone DA, Giesler GJ., Jr. Relief of itch by scratching: state-dependent inhibition of primate spinothalamic tract neurons. Nat. Neurosci. 2009;12:544–546. [PMC free article] [PubMed] [Google Scholar]
63. Papoiu AD, Coghill RC, Kraft RA, Wang H, Yosipovitch G. A tale of two itches. Common features and notable differences in brain activation evoked by cowhage and histamine induced itch. Neuroimage. 2012;59:3611–3623. [PMC free article] [PubMed] [Google Scholar]
64. Liu B, et al. TRPA1 controls inflammation and pruritogen responses in allergic contact dermatitis. FASEB J. 2013;27:3549–3563. [PMC free article] [PubMed] [Google Scholar]
65. Maurer M, et al. Omalizumab for the treatment of chronic idiopathic or spontaneous urticaria. N. Engl. J. Med. 2013;368:924–935. [PubMed] [Google Scholar]
66. Grimstad O, et al. Anti-interleukin-31-antibodies ameliorate scratching behaviour in NC/Nga mice: a model of atopic dermatitis. Exp. Dermatol. 2009;18:35–43. [PubMed] [Google Scholar]
67. Basbaum AI, Bautista DM, Scherrer G, Julius D. Cellular and molecular mechanisms of pain. Cell. 2009;139:267–284. [PMC free article] [PubMed] [Google Scholar]
68. Tani E, Ishikawa T. Histamine acts directly on calcitonin gene-related peptide-and substance P-containing trigeminal ganglion neurons as assessed by calcium influx and immunocytochemistry. Auris Nasus Larynx. 1990;17:267–274. [PubMed] [Google Scholar]
69. Zhou Z, Davar G, Strichartz G. Endothelin-1 (ET-1) selectively enhances the activation gating of slowly inactivating tetrodotoxin-resistant sodium currents in rat sensory neurons: a mechanism for the pain-inducing actions of ET-1. J. Neurosci. 2002;22:6325–6330. [PMC free article] [PubMed] [Google Scholar]
70. Gomes LO, Hara DB, Rae GA. Endothelin-1 induces itch and pain in the mouse cheek model. Life Sci. 2012;91:628–633. [PubMed] [Google Scholar]
71. Liang J, Ji Q, Ji W. Role of transient receptor potential ankyrin subfamily member 1 in pruritus induced by endothelin-1. Neurosci. Lett. 2011;492:175–178. [PubMed] [Google Scholar]
72. Kim DK, et al. Involvement of serotonin receptors 5-HT1 and 5-HT2 in 12(S)-HPETE-induced scratching in mice. Eur. J. Pharmacol. 2008;579:390–394. [PubMed] [Google Scholar]
73. Stefansson K, et al. Activation of proteinase-activated receptor-2 by human kallikrein-related peptidases. J. Invest. Dermatol. 2008;128:18–25. [PubMed] [Google Scholar]
74. Tsujii K, Andoh T, Lee JB, Kuraishi Y. Activation of proteinase-activated receptors induces itch-associated response through histamine-dependent and -independent pathways in mice. J. Pharmacol. Sci. 2008;108:385–388. [PubMed] [Google Scholar]
75. Zhu Y, et al. Participation of proteinase-activated receptor-2 in passive cutaneous anaphylaxis-induced scratching behavior and the inhibitory effect of tacrolimus. Biol. Pharm. Bull. 2009;32:1173–1176. [PubMed] [Google Scholar]
76. Bautista DM, Pellegrino M, Tsunozaki M. TRPA1: a gatekeeper for inflammation. Annu. Rev. Physiol. 2013;75:181–200. [PMC free article] [PubMed] [Google Scholar]
77. Tominaga M, Caterina MJ. Thermosensation and pain. J. Neurobiol. 2004;61:3–12. [PubMed] [Google Scholar]
78. Miyamoto T, Nojima H, Shinkado T, Nakahashi T, Kuraishi Y. Itch-associated response induced by experimental dry skin in mice. Jpn. J. Pharmacol. 2002;88:285–292. [PubMed] [Google Scholar]
79. Akiyama T, Carstens MI, Carstens E. Spontaneous itch in the absence of hyperalgesia in a mouse hindpaw dry skin model. Neurosci. Lett. 2010;484:62–65. [PMC free article] [PubMed] [Google Scholar]
80. Jin H, He R, Oyoshi M, Geha RS. Animal models of atopic dermatitis. J. Invest. Dermatol. 2009;129:31–40. [PMC free article] [PubMed] [Google Scholar]
81. Seike M, Ikeda M, Kodama H, Terui T, Ohtsu H. Inhibition of scratching behaviour caused by contact dermatitis in histidine decarboxylase gene knockout mice. Exp. Dermatol. 2005;14:169–175. [PubMed] [Google Scholar]
82. Dey DD, Landrum O, Oaklander AL. Central neuropathic itch from spinal-cord cavernous hemangioma: a human case, a possible animal model, and hypotheses about pathogenesis. Pain. 2005;113:233–237. [PubMed] [Google Scholar]
83. D'Almeida JA, et al. Behavioral changes of Wistar rats with experimentally-induced painful diabetic neuropathy. Arq. Neuropsiquiatr. 1999;57:746–752. [PubMed] [Google Scholar]
84. Stebbins KJ, et al. Therapeutic efficacy of AM156, a novel prostanoid DP2 receptor antagonist, in murine models of allergic rhinitis and house dust mite-induced pulmonary inflammation. Eur. J. Pharmacol. 2010;638:142–149. [PubMed] [Google Scholar]
85. Zheng T, et al. Transgenic expression of interleukin-13 in the skin induces a pruritic dermatitis and skin remodeling. J. Invest. Dermatol. 2009;129:742–751. [PMC free article] [PubMed] [Google Scholar]
86. Hansson L, et al. Epidermal overexpression of stratum corneum chymotryptic enzyme in mice: a model for chronic itchy dermatitis. J. Invest. Dermatol. 2002;118:444–449. [PubMed] [Google Scholar]
87. Konishi H, et al. IL-18 contributes to the spontaneous development of atopic dermatitis-like inflammatory skin lesion independently of IgE/stat6 under specific pathogen-free conditions. Proc. Natl. Acad. Sci. USA. 2002;99:11340–11345. [PMC free article] [PubMed] [Google Scholar]
88. Briot A, et al. Par2 inactivation inhibits early production of TSLP, but not cutaneous inflammation, in Netherton syndrome adult mouse model. J. Invest. Dermatol. 2010;130:2736–2742. [PubMed] [Google Scholar]
89. Brandt EB, Gibson AM, Bass S, Rydyznski C, Khurana Hershey GK. Exacerbation of allergen-induced eczema in TLR4- and TRIF-deficient mice. J. Immunol. 2013;191:3519–3525. [PMC free article] [PubMed] [Google Scholar]
90. Turner MJ, et al. Topical application of a vitamin D analogue exacerbates atopic dermatitis and induces the atopic dermatitis-like phenotype in Stat6VT mice. Pediatr. Dermatol. 2013;30:574–578. [PMC free article] [PubMed] [Google Scholar]
91. Nakajima K, et al. Barrier abnormality due to ceramide deficiency leads to psoriasiform inflammation in a mouse model. J. Invest. Dermatol. 2013;133:2555–2565. [PubMed] [Google Scholar]
92. Yockey LJ, et al. The absence of a microbiota enhances TSLP expression in mice with defective skin barrier but does not affect the severity of their allergic inflammation. J. Invest. Dermatol. 2013;133:2714–2721. [PMC free article] [PubMed] [Google Scholar]
93. Yoo J, et al. Spontaneous atopic dermatitis in mice expressing an inducible thymic stromal lymphopoietin transgene specifically in the skin. J. Exp. Med. 2005;202:541–549. [PMC free article] [PubMed] [Google Scholar]
94. Oh MH, et al. TRPA1-dependent pruritus in IL-13-induced chronic atopic dermatitis. J. Immunol. 2013;191:5371–5382. [PMC free article] [PubMed] [Google Scholar]
95. Wallace K, et al. The PXR is a drug target for chronic inflammatory liver disease. J. Steroid Biochem. Mol. Biol. 2010;120:137–148. [PMC free article] [PubMed] [Google Scholar]
96. Belghiti M, et al. Potentiation of the transient receptor potential vanilloid 1 channel contributes to pruritogenesis in a rat model of liver disease. J. Biol. Chem. 2013;288:9675–9685. [PMC free article] [PubMed] [Google Scholar]
97. Yilinuer H, Yamaoka J, Kawana S. Effect of epinastine hydrochloride on murine self-scratching behavior after skin-scratching stimulation. Arch. Dermatol. Res. 2010;302:19–26. [PubMed] [Google Scholar]
98. Honma Y, et al. Cyclooxygenase-1 inhibition delays recovery of the cutaneous barrier disruption caused by mechanical scratching in mice. Br. J. Dermatol. 2007;156:1178–1187. [PubMed] [Google Scholar]