U.S. flag

An official website of the United States government

NCBI Bookshelf. A service of the National Library of Medicine, National Institutes of Health.

Emir TLR, editor. Neurobiology of TRP Channels. Boca Raton (FL): CRC Press/Taylor & Francis; 2017. doi: 10.4324/9781315152837-13

Cover of Neurobiology of TRP Channels

Neurobiology of TRP Channels.

Show details

Chapter 13Airway Pathogenesis Is Linked to TRP Channels

.

13.1. Introduction

Respiratory diseases affect the quality of life of an important population worldwide. In light of this, efforts have been made to determine which ion channels are involved in diseases including asthma, chronic obstructive pulmonary disease (COPD), chronic cough, pulmonary arterial hypertension, cystic fibrosis, allergic and nonallergic rhinitis, acute lung injury, and idiopathic pulmonary fibrosis. In the context of some of these diseases, it has been found that the function and/or the expression of some TRP channels may be altered. Among these the TRPA1, TRPV1, TRPV4, and TRM8 channels have been linked to asthma, COPD, and chronic cough among other maladies. This chapter focuses on the role of these channels in normal and abnormal airway function.

13.2. Normal airway physiology of TRP channels

The sensory nerve innervation of the respiratory system is relatively well understood and includes innervation of the smooth muscle and glands of the larynx, trachea, bronchial tree, and lungs (Widdicombe, 1995; Canning, 2011). When activated, sensory afferent nerve impulses are carried via the vagus nerve to the central nervous system, leading to central reflexes including cough, dyspnea, and changes in breathing pattern. Different subtypes of sensory afferent fibers are known to exist including rapidly adapting receptors (RARs), slowly adapting receptors (SARs), and C-fibers (Canning et al., 2006). Two subpopulations of C-fibers are derived from cell bodies found in the jugular ganglion and nodose ganglion. The C-fibers from nodose neurons primarily innervate intrapulmonary structures, whereas jugular ganglia neurons project to the trachea, extrapulmonary bronchi, and lung paranchymal tissue (Spina et al., 2013).

Sensory neurons in the upper airway are thought to protect the lower airway by limiting exposure to foreign particles including environmental irritants and noxious chemicals. This is achieved by pathways inducing inflammation, mucus secretion, airway constriction, and reflexes such as cough and sneezing. There is emerging evidence to suggest that TRP channels may play an important role in these airway protective mechanisms. TRP channels have been identified in neuronal and nonneuronal cells of the lung including sensory afferents, epithelia, airway smooth muscle (ASM), immune cells, and inflammatory cells (Nassini et al., 2012) (Figure 13.1). The TRP channels that have been looked at in depth and are discussed in this chapter are TRPA1, TRPV1, TRPV4, and TRPM8.

Figure 13.1. Localization and function of TRP channels in neuronal and nonneuronal cells of the respiratory tract.

Figure 13.1

Localization and function of TRP channels in neuronal and nonneuronal cells of the respiratory tract. (RARs, rapidly adapting receptors; SARs, slowly adapting receptors.)

13.2.1. TRPA1

Transient receptor potential ankyrin 1 (TRPA1) is one of the most extensively studied channels in the airway, and emerging evidence points toward a major role for TRPA1 in the lung's defense system. Under normal conditions, the main functions of TRPA1 are protecting the airway by evoking cough associated with respiratory irritants and minimizing exposure by controlling airway tone causing bronchoconstriction.

Much of the evidence comes from electrophysiology and calcium imaging experiments and utilizing TRPA1 knockout mice and selective TRPA1 antagonists. TRPA1 is referred to as an irritant receptor as it can be activated by exogenous irritants including mustard oil, wasabi, cinnamaldehyde, cigarette smoke, chlorine, aldehydes, and scents (Caceres et al., 2009; Jordt et al., 2004). Activation by endogenous stimuli has also been identified, including reactive oxygen species, hypochlorites, lipid peroxidation products, isoprostanes, and prostaglandins (Bandell et al., 2004; Jha et al., 2015; Bautista, et al., 2005; Andersson et al., 2008).

Quantitative analysis from mouse tissue has shown that TRPA1 channels are mainly expressed in the spinal dorsal root ganglia (DRG) and trigeminal ganglia (TG) (Jang et al., 2012). Expression has also been demonstrated in mouse bronchopulmonary afferent neurons (jugular/nodose) with apparent coexpression of TRPA1 and TRPV1 (Story et al., 2003; Nassenstein et al., 2008). As well as neuronal tissue, evidence also points toward the presence of TRPA1 channels in nonneuronal cells. Using immunohistochemistry, TRPA1 staining has been observed in cultured human and mouse airway epithelial cells and smooth muscle cells (Nassini, 2012). Functional expression has also been demonstrated in human lung fibroblast cells (CCD19-Lu) and in the human pulmonary alveolar epithelial cell line (A549) at both mRNA and protein levels (Mukhopadhyay et al., 2011; Büch et al., 2013).

Activation of TRPA1 leads to neuronal depolarisation and action potential discharge in the afferent fiber. Exogenous compounds such as mustard oil and cinnamaldehyde have been shown to activate TRPA1 channels directly, by covalent modification of reactive cysteine residues in the amino terminus of the protein (Macpherson, 2007; Hinman, 2006). Inflammatory mediators such as bradykinin activate TRPA1 indirectly via G protein-coupled receptors. This leads to a second messenger cascade and increase in intracellular Ca2+ (Bandell, 2004). Several studies have demonstrated the importance of Ca2+ in the regulation of TRPA1 activity; intracellular Ca2+ release may directly activate cation influx through TRPA1 channels (Zurgorg, 2007).

Recent data have demonstrated the importance of the TRPA1 pathway in initiating the normal cough reflex (Birrell et al., 2009; Andrè et al., 2009; Brozmanova et al., 2012; Grace et al., 2012). Isolated vagal nerve preparations have shown that tussive agents activate vagal nerve endings in mouse, guinea pig, and human tissue. A role for TRAP1 was confirmed using TRPA1−/– mice and TRPA1 inhibitors. Furthermore, cough evoked in an in vivo guinea pig model and human volunteers was shown to be reduced in the presence of TRPA1 antagonists (Birrell et al., 2009). Studies have identified several triggers of the TRPA1-mediated cough response including exogenous irritants such as those present in cigarette smoke, agents produced endogenously by oxidative stress (Birrell et al., 2009; Andrè et al., 2009; Brozmanova et al., 2012), and the endogenous inflammatory mediators prostaglandin E2 and bradykinin (Grace et al., 2012). As such endogenous mediators are known to be elevated in certain pathological conditions, these findings have implications for the role of TRPA1 channels in disease-associated cough. This is discussed along with the role of pollution in Section 13.3.

A further protective role for TRAP1 has been reported by the modulation of airway tone. Components of cigarette smoke have been shown to cause contraction of isolated guinea pig bronchial rings by activation of TRPA1 channels on sensory neurons. Channel activation triggers tachykinin release, which in turn acts on smooth muscle cells to initiate contraction, a protective mechanism, referred to as the nocifensor system (Andrè, 2008). In addition, the TRPA1 agonist 4-oxo-2-nonenal (4-ONE) and the general anesthetics isoflurane and desflurane have been shown to induce TRPA1-dependent bronchial contraction by a similar mechanism (Satoh et al., 2009). In a more recent study, acrolein a chemical component of smoke, has been shown to relax isolated tracheal smooth muscle (Cheah et al., 2014). Authors suggest relaxation may serve as a defensive mechanism by counteracting the effect of uncontrolled neurogenic excitation and airway narrowing by irritants evoking inflammation and bronchoconstriction. This effect is thought to occur via a pathway involving cross-talk between TRPA1 expressing sensory C-fibers, epithelial cells, and smooth muscle (Cheah et al., 2014). TRPA1 channels have also been implicated in the release of proinflammatory mediators from nonneuronal cells including epithelium, smooth muscle cells, and fibroblasts (Grace et al., 2014).

13.2.2. TRPV1

TRPV1-expressing neurons represent the largest subpopulation of primary sensory neurons. TRPV1-positive neurons belong to cells with C- and Aδ- fibers and are functionally identified as nociceptors. They innervate the entire respiratory tract from the nose to the alveoli including smooth muscle and blood vessels (Watanabe et al., 2006; Grace et al., 2014). TRPA1-positive neurons are most notably stretch insensitive and peptidergic, and are coexpressed with TRPV1 channels. However, TRPV1-postive neurons can be both stretch insensitive and stretch sensitive and either peptidergic or nonpeptidergic. They are also not exclusively expressed with TRPA1 (Bhattacharya et al., 2008).

Observations have reported a positive interaction between TRPA1 and TRPV1 regulating the sensitivity of sensory neurons during an inflammatory reaction (Lee and Hsu, 2015). Like TRPA1, TRPV1 stimulation may also be the consequence of G protein-coupled receptor (GPCR) activation by inflammatory mediators, such as bradykinin (Bautista et al., 2006; Grace et al., 2012). Low levels of TRPV1 in the airway may suggest that TRPV1 has limited physiological function, but its ability to be activated by proinflammatory mediators suggests that this channel may play an important role in inflammatory conditions and be of pathological importance. As is the case with TRPA1, TRPV1 has also been implicated in playing a role in airway defense, and several studies have been carried out investigating the role of TRPV1 sensory nerves in the initiation of the normal cough reflex. Cough induced in human and animal models by substances that activate TRPV1 such as capsaicin and citric acid is inhibited by TRPV1 antagonists (Bhattacharya et al., 2007; Grace et al., 2014). Investigators have also proposed that TRPV1 is involved in modulation of airway tone, and bronchoconstriction occurs by a mechanism involving activation of TRPV1 by inflammatory mediators (Delescluse et al., 2012). In this study, increased airway contraction in a disease model was inhibited by TRPV1 antagonists. Further evidence for a role in defense mechanisms has been shown by a study demonstrating that environmental prototype particulate matter (PM) is sensed by TRPV1 (Deering-Rice et al., 2012). This study showed that TRPV1 mediates the induction of several important proinflammatory cytokine/chemokine genes in lung epithelial cells in response to PM.

It is also important to note the presence and function of TRPV1 in nonneuronal cells in the airway. Functional TRPV1 protein has been demonstrated in cultured primary bronchial epithelial cells by patch-clamp experiments (McGarvey et al., 2014). Stimulation by capsaicin induced a dose-dependent release of interleukin 8 (IL-8), which, being blocked by capsazepine, suggests TRPV1 activation (McGarvey et al., 2014). TRPV1 has also been shown to be expressed in smooth muscle cells and to play a role in airway smooth muscle physiology, including proliferation and apoptosis (Zhao, 2013) and control of airway tone (Grace et al., 2014). Several recent studies have demonstrated expression of TRP channels in inflammatory and immune cells (Parenti et al., 2016). As discussed later in the chapter, emerging evidence suggests that TRP channels may be responsible at least in part for the transition of early defensive immune and inflammatory responses to chronic responses and disease pathology (Parenti et al., 2016). In relation to TRPV1, murine CD4+ T lymphocytes have been shown to express functional TRPV1, which was activated on stimulation of the T cell antigen receptor (TCR), contributed to Ca2+ influx and TCR signaling, resulting in T-cell activation. Inhibition of TRPV1 in mouse and human CD4+ T cells, with antagonists or by genetic manipulation, resulted in a Trpv1−/− CD4+ T-cell-like phenotype (Bertin et al., 2014).

13.2.3. TRPV4

Compared to TRPA1 and TRPV1, little is known about TRPV4 expression in peripheral sensory neurons that innervate the lung. However, a recent study identifies the TRPV4-ATP-P2X3 interaction as a key osmosensing pathway involved in airway sensory nerve reflexes (Bonvini, 2016). TRPV4 ligands and hypoosmotic solutions caused depolarization of murine, guinea pig, and human vagus nerves, and firing of Aδ-fibers, which was inhibited by TRPV4 and P2X3 receptor antagonists; both antagonists blocked TRPV4-induced cough. In another study, TRPV4 has also been implicated in the development of neurogenic inflammation (Vergnolle et al., 2010). Hypotonic solutions and 4αPDD have been shown to induce neuropeptide release from afferent nerves in isolated murine airways (Vergnolle et al., 2010). These data suggest that TRPV4 also plays a role in airway defense.

Several studies have demonstrated that TRPV4 is widely expressed in nonneuronal cells in the airways, including mRNA expression in human airway smooth muscle cells (Jia et al., 2004; Dietrich et al., 2006), confirmation of expression by immunohistochemistry in the alveolar septal wall of human, rat and mouse lung (Alvarez et al., 2006), and human nasal, tracheal and bronchial epithelial cells (Alenmyr et al., 2014; Fernández-Fernández et al., 2008). TRPV4 is also present in lung fibroblasts (Rahaman et al., 2014) and inflammatory cells including alveolar macrophages (Hamanaka et al., 2010) and mononuclear cells (Delany et al., 2001).

Compared to other TRP channels, less is also known about external and endogenous ligands that may be responsible for TRPV4 activation and function, and indeed the mechanisms involved. However, known channel activators include osmotic changes, for example, increasing activity in hypotonic solutions, and mechanical stimuli such as shear stress (Garcia-Elias et al., 2014). Both stimuli depend on phospholipase A2 activation and subsequent production of arachidonic acid (AA) metabolites. TRPV4 channels can also be directly activated by AA (Zheng et al., 2013). Other TRPV4 channel activators include moderate heat (24°C–38°C) and the synthetic phorbol ester 4α-PDD (Watanabe et al., 2002a). Evidence suggests that TRPV4 channel activity is regulated by Ca2+, depending on its concentration. Ca2+ can either potentiate or inhibit channel activity (Garcia-Elias et al., 2014).

A role for TRPV4 in airway smooth muscle function has been demonstrated in human bronchial smooth muscle cells, where 4α-PDD, stretch, and hypotonic solutions caused increases in Ca2+ and subsequent contraction (Jia et al., 2004). An indirect role for TRPV4 in contraction has been shown in human bronchial smooth muscle segments using the TRPV4 agonist GSK1016790 (McAlexander et al., 2014). This agonist was shown to cause contraction via 5-lipoxygenase activation and production of cysteinyl leukotrienes (McAlexander et al., 2014).

Further evidence for the involvement of TRPV4 in lung defense has been shown by studies demonstrating activation of TRP4 modulating ciliary beat frequency (Lorenzo et al., 2008) and the control of epithelial and endothelial barrier function (Alvarez et al., 2006; Li et al., 2011a). A potential role in macrophage activation by mechanical stress (Hamanaka et al., 2010) has also been suggested.

13.2.4. TRPM8

Evidence to date suggests TRPM8 may play a minimal role in airway physiology with contradictory findings often reported (Grace et al., 2014). One study demonstrated that TRPM8 is not expressed in mouse vagal afferents innervating the airways, and menthol, a TRPM8 agonist, was ineffective at increasing intracellular Ca2+ in lung-specific vagal sensory neurons (Nassenstein et al., 2008). Other studies have found evidence of a subpopulation of rat airway vagal afferent nerves expressing TRPM8 receptors and that these receptors are activated by cold temperatures (Xing et al., 2008). It is implied from these data that activation is likely to trigger responses such as airway constriction in response to a reduction in temperature. Molecular analysis has also identified expression of TRPM8 in a subset of nasal trigeminal afferent neurons (Plevkova et al., 2013). However, authors have suggested that activation of TRPM8 has an inhibitory effect; they conclude that menthol suppresses cough evoked in the lower airways primarily through this subset of neurons initiated from the nose (Buday et al., 2012; Plevkova et al., 2013).

TRPM8 is also expressed in airway epithelial cells and has recently been shown to inhibit proliferation and migration in a rat asthma model (Zhang, 2016). TRPM8 has been identified primarily within endoplasmic reticulum membranes of epithelial cells (Sabnis et al., 2008) and may play a role in production of mucus and inflammatory mediators from airway epithelium (Sabnis et al., 2008; Grace et al., 2014).

13.3. TRP channels and airway pathophysiology

As already highlighted, TRP channels play a definitive role in protecting the airway from foreign particles. When activated, TRP channels can decrease respiratory drive, trigger cough, induce airway narrowing by modulating airway tone, and induce a coordinated inflammatory response. Evidence is emerging for the involvement of TRP channels in the cross-talk between immunogenic and neurogenic pathways in airway inflammation, as these channels play a key role in the response of sensory neurons to inflammatory mediators. When exaggerated, these same mechanisms may give rise to neurogenic inflammation, airway hyperactivity (AHR), and abnormal cough. Such symptoms are associated with diseases such as asthma, COPD, and chronic cough.

13.3.1. Asthma

Asthma is a chronic inflammatory airway disease affecting an estimated 300 million people worldwide. The number of annual deaths attributed to the disease is approximately 250,000. Physiological hallmarks of asthma include AHR and variable airflow obstruction. The disease is defined by symptoms including wheeze, chest tightness, shortness of breath, cough, and dyspnea. The airway inflammatory response can vary between asthmatic patients, and as the disease becomes more persistent, further airway obstruction can occur from edema, mucus hypersecretion, and airway remodeling. Repeated exposure to environmental allergens is thought to be the underlying cause of asthma, the early asthmatic response occurring minutes after initial exposure to allergen. Recognition of the allergen by IgE on the surface of mast cells induces degranulation and release of inflammatory mediators such as histamine and cysteinyl leukotriene (CysLT), causing acute bronchoconstriction (Adelroth et al., 1986). A late asthmatic response (LAR) can follow 3–8 hours after allergen exposure and only occurs in approximately 50% of patients (O'Byrne et al., 2009). The late asthmatic response results in further bronchoconstriction and is also associated with a complex immune and inflammatory response. This response is linked to an increase in inflammatory cells including Th-2 cells, eosinophils, airway, basophils, and less consistently neutrophils and trafficking and activation of myeloid dendritic cells into the airways (Holgate et al., 2008; Gauvreau et al., 2015). Prominent inflammatory mediators include IL-4, IL-5, IL-13, eotaxin (CCL11), and eicosanoids. The LAR phenotype can also occur in response to other triggers such as exercise, cold air, irritants, and stress. The mechanisms regulating the airway response to these factors are less well defined. Isocyanate is another known trigger in patients suffering from occupational asthma (Kenyon et al., 2012). Disease exacerbations in asthmatic patients can also be caused by respiratory infections.

TRP channels are likely candidates in the pathophysiology of asthma, as several exogenous and endogenous TRP channel agonists such as cigarette smoke and reactive oxygen species are also known asthma triggers. Irritants can trigger asthma-type symptoms and subsequent hypersensitivity to chemical and physical stimuli (Preti et al., 2012). Endogenous mediators produced by infiltrating immune cells or inflamed airway tissue can reach elevated levels high enough to chronically activate TRPA1 in airway sensory neurons (Trevisani et al., 2007). A critical role for TRPA1 in the pathogenesis of asthma has been shown in several studies using the TRPA1 knockout mouse. These mice have been shown to be deficient in the neuronal detection of proinflammatory asthma agents (Bessac et al., 2008; Andrè et al., 2008). Lack of TRPA1 may prevent neuronal excitation and Ca2+ influx normally activated by these mediators, which is likely to occur during inflammatory progression in asthma (Caceres et al., 2009). Further evidence of a role for TRPA1 in asthmatic airway inflammation has been shown in the mouse ovalbumin model of asthma (Caceres et al., 2009). In this study, pharmacological inhibition and genetic deletion of TRPA1 diminish allergen-induced inflammatory leukocyte infiltration, mucus production, cytokine and chemokine levels, and AHR. TRPAP1 knockout mice also show impaired acute and inflammatory neuropeptide release in the airways required for leukocyte infiltration. A role for TRPV1 is less clear as data from asthma models have reported conflicting results since lack of TRPV1 did not demonstrate any change in the mouse ovalbumin model (Caceres et al., 2009). However, other asthma models have shown that modulation of TRPV1 attenuated the asthma model inflammatory phenotype (Mori et al., 2011; Rehman et al., 2013).

Evidence of a role for TRP channels in control of airway tone was discussed earlier in the chapter. It is therefore not surprising to find studies linking TRP channel activation and AHR in asthma models. The results of one such study using ovalbumin-sensitized animals, show abolished LAR response in the presence of TRP channel inhibitors (Raemdonck et al., 2012). The authors conclude that allergen challenge activates TRPA1 channels on sensory neurons, which in turn triggers a central reflex leading to parasympathetic cholinergic activation of airway smooth muscle. Although this study ruled out a role for TRPV1 in AHR, data from unanesthetized, ovalbumin-sensitized guinea pig showed inhibition of AHR by TRPV1 antagonists (Delescluse et al., 2012). A recent study has demonstrated a link between TRPA1 and TRPV1 in the development of AHR in response to the chemical sensitizer toluene-2,4-diisocyanate (TDI) (Devos et al., 2016). Authors used a mouse model of TDI sensitization to induce AHR, and blocked TRPA1- and TRPV1-channel activity using pharmacological methods and knockout mice. They put forward a neuroimmune pathway as a possible mechanism after demonstrating that TRPA1 and TRPV1 channels and mast cell knockout mice did not exhibit AHR despite being sensitized by TDI.

The discovery of genetic variants of certain TRP channels has led to studies demonstrating a link between the presence of certain polymorphisms and altered disease presentation. The outcome of one such study showed an association between a TRPV1 single nucleotide polymorphism (SNP), and a protective effect against the presence of wheezing in a group of asthma patients (Cantero-Recasens et al., 2010). Fluorescent microscopy revealed a lower increase in Ca2+ concentration upon stimulation indicating a decrease in channel activity. These results provide evidence for a role of TRPV1 channels in altered Ca2+ signaling underlying asthma pathophysiology. Another study has implicated TRPA1 in the modulation of asthma in children exposed to high levels of pollution (Deering-Rice et al., 2015). Expression of variant forms of TRPA1 may increase the sensitivity of TRPA1 to insoluble particulate matter including 3,5-ditert butylphenol, a soluble nonelectrophilic agonist and component of diesel exhaust particles. Such polymorphisms of the TRPA1 channel may lead to an increase in activation and a correlation with reduced asthma control.

As the role of TRP channels in asthma pathophysiology becomes more apparent, changes in expression levels or patterns of TRP channel expression become increasingly likely. A search of the literature revealed a study demonstrating altered TRPV1 expression in a rat model of chronic inflammation induced by an ovalbumin sensitization (Zhang et al., 2008). Authors report an increase in the proportion of TRPV1 expressing neurons in pulmonary myelinated afferents in the nodose ganglia. Data also showed an increase in sensitivity to capsaicin of vagal bronchopulmonary myelinated afferents. Further studies using a similar model in guinea pigs have shown that inflammation causes a “phenotypic switch in vagal tracheal cough-causing, low-threshold mechanosensitive Aδ neurons, such that they begin expressing TRPV1 channels” (Lieu et al., 2012). Authors also provide evidence to suggest a role for neurotrophic factors in the modulation of gene expression of TRPV1 and nerve phenotype, which could contribute to excessive coughing caused by allergens. Other results have shown an increase in sensitivity to continuous capsaicin inhalation in cough-variant asthma (Nakajima et al., 2006). Other evidence has highlighted the role of TRPV1 and TRPA1 in sensitization resulting in exaggerated sensory responses to ROS causing airway hypersensitivity—a characteristic feature of asthma (Ruan et al., 2014).

13.3.2. Chronic obstructive pulmonary disease

Chronic obstructive pulmonary disease (COPD) is a term used to describe a number of conditions including emphysema that affects the alveoli, and chronic bronchitis affecting the bronchi. According to the Global Initiative for COPD guidelines, it is characterized by persistent airflow limitation that is usually progressive and associated with an enhanced chronic inflammatory response to noxious particles or gases (Pauwels et al., 2001). It is well established that tobacco smoke increases susceptibility to COPD, but other factors including genetic factors and prolonged exposure to occupational dusts and chemicals can also increase the risk of developing the disease (Salvi and Barnes, 2009). Symptoms include chronic cough, dyspnea, and excess sputum production; exacerbations also often occur. Examination of smokers with chronic bronchitis has shown an increased number of CD8+ T lymphocytes, neutrophils, and macrophages in surgical pulmonary specimens (Saetta et al., 1997; Saetta et al., 1998). Increased levels of proinflammatory mediators have also been shown in patients with COPD, which are significantly associated with disease severity (Barnes, 2016; Selvarajah et al., 2016).

Recent evidence suggests that TRPA1 is the primary TRP channel implicated in the pathogenesis of COPD. As described earlier, it is activated by several components of cigarette smoke, including acrolein and crotonaldehyde and nicotine (Lin et al., 2010; Andrè et al., 2008; Talavera et al., 2009), so a role of TRPA1 is highly likely. Cigarette smoke aqueous extract (CSE) has been shown to mobilize Ca2+ in cultured guinea pig jugular ganglia neurons and promote contraction of isolated guinea pig bronchi (Andrè et al., 2008). CSE and unsaturated aldehydes contract the guinea pig bronchus by a neurogenic mechanism mediated by TRPA1 stimulation, Ca2+ mobilization, and release of neuropeptides. Another study by Shapiro and colleagues has investigated the effects of wood smoke particle material (WSPM), which has been shown to have a clear impact on human health causing a progressive decline in lung function and development of chronic COPD (Naeher et al., 2007; Laumbach et al., 2012; Shapiro et al., 2013). This study demonstrated that TRPA1 in trigeminal sensory neurons was activated by WSPM resulting in Ca2+ flux shown by imaging experiments. Continued exposure to WSPM therefore has implications for the development of neurogenic inflammation, decreased respiratory drive, chronic cough, and bronchoconstriction by activating this pathway (Shapiro et al., 2013).

There is emerging evidence to suggest the involvement of other TRP channels in respiratory symptoms associated with COPD. The suggestion of a possible role for TRPV1 comes from the effects of tiotropium, a widely prescribed drug for its bronchodilator effects in COPD and asthma. It has also been shown to attenuate cough in preclinical and tussive challenge studies (Bateman et al., 2009). Recent data have shown that tiotropium inhibited capsaicin-induced Ca2+ responses and currents in isolated guinea pig vagal tissue, implying a TRPV1 neuronal mediated effect (Birrell et al., 2014). Increased TRPV1 and TRPV4 mRNA expression has also been demonstrated in patients with COPD (Baxter et al., 2014). In the same study, TRPV1 and TRPV4 were shown to play a role in cigarette smoke (CS)–induced release of extracellular ATP, which is elevated in the COPD airway and is thought to be intrinsic to CS-induced inflammation. Moreover, SNPs in TRVP4 have been associated with COPD (Zhu et al., 2009).

13.3.3. Chronic cough

Chronic cough is typically defined as a cough that lasts longer than 8 weeks and occurs in approximately 40% of the population. As discussed above, chronic cough is present in diseases such as COPD and asthma, but can also occur as a persistent cough after a viral or bacterial infection or in isolation with an unknown cause. The underlying mechanisms of chronic cough are complex, and causes of exaggerated cough in disease pathology are not completely understood. As already described, TRP channels play an important role in the cough reflex as a protective function in airway defense so are likely to be involved in abnormal cough presentation. A role for TRPV1 in chronic cough is a likely scenario due to the protussive effects of the TRPV1 agonist capsaicin. Increased expression of TRPV1 in sensory nerves has been demonstrated in patients with chronic cough (Mitchell et al., 2005; Groneberg et al., 2004). Moreover, in a recent study an association has been made between the presence of TRPV1 SNPs and a higher risk of developing chronic cough in patients who smoke, or with occupational exposure (Smit et al., 2012). Studies have also shown an increase in sensitivity to capsaicin in patients with chronic cough (Doherty et al., 2000; Pecova et al., 2008). The TRPA1 channel is also involved in cough associated with asthma and COPD as described above. TRPA1 is activated by reactive electrophilic molecules including acrolein that is a by-product of oxidative stress that can lead to activation of the cough reflex (Bautista et al., 2006; Trevisani et al., 2007). As oxidative stress can be initiated by environmental irritants and as a by-product of inflammation, this is a possible pathway whereby exaggerated cough is initiated. This is made more likely during inflammation by the increase of other endogenous ligands that directly activate TPA1, such as prostaglandins (Grace and Belvisi, 2011).

Patients suffering from idiopathic cough have been reported to have suffered from a viral infection preceding the onset of their cough (Haque et al., 2005). Human rhinovirus is the major cause of the common cold and has also been shown to be responsible for asthma exacerbations (Arden et al., 2006; Nicholson et al., 1993). A recent study has demonstrated the ability of rhinovirus to infect neuronal cells, and that infection causes an increase in expression of TRPV1, TRPA1, and TRPM8 mRNA and protein (Abdullah et al., 2014). Virus-induced soluble factors, for example, IL-8, IL-16, and nerve growth factor, were shown to be sufficient for TRPV1 and TRPA1 upregulation in culture, Interestingly, mechanisms of upregulation differed for TRPM8 which required replicating virus. This suggests that direct interaction with virus particles and neuronal cells is required for TRPM8 upregulation (Abdullah et al., 2014). Taken together these data suggest a role for certain TRP channels on sensory afferents in postviral cough.

13.3.4. Other respiratory diseases

Nonallergic rhinitis (NAR) is inflammation of the nose that affects approximately 10% of people worldwide. Symptoms include sneezing, irritation, and nasal congestion that are induced by hypersensitivity to stimuli including smoke, temperature changes, and irritants. TRPV1 channels on C-fiber afferents innervating the nasal mucosa are thought to play a role in NAR, since the activation of this ion channel causes the release of neuropeptides substance P and calcitonin G-related peptides from nerve terminals resulting in a local inflammatory response (Van Gerven et al., 2012). Nasal application of capsaicin has been shown to alleviate the symptoms of TRPV1. This is thought to occur by desensitization after strong excitation of the TRPV1 neurons, or massive influx of Ca2+ resulting in nonfunctional afferents or degeneration of nerve terminals (Anand and Bley, 2011; Van Gerven et al., 2014). The role of neuronal TRP channels in allergic rhinitis is less clear. It has been hypothesized that allergic rhinitis induced by exposure to allergens may have increased sensitivity to stimulation of TRP channels (Alenmyr, 2009). This study went on to demonstrate an increase in itch response to stimulation of TRPV1 at seasonal allergen exposure. However, the TRPV1 blocker SB-705498 had no effect on nasal symptoms of patients with seasonal allergic rhinitis including itch (Alenmyr et al., 2012). More recent studies have concentrated on alternative target cells, demonstrating localization of TRPV1 expression to CD4+ T cells (Samivel et al., 2016). TRPV1 has been shown to be involved in T-cell receptor signaling that is altered in TRPV1 knockout allergic rhinitis mice (Samivel et al., 2016).

Apneic responses (lethal ventilator arrest) have recently been shown to be associated with upregulation of TRPV1 channels (Zhuang et al., 2015). Prenatal exposure to nicotine has been shown to trigger apneic responses during severe hypoxia (Zhuang et al., 2014). This is caused by lack of inspiratory drive from the CNS but the underlying mechanisms are poorly understood. The study by Zhuang et al. (2015) has demonstrated an association between apneic responses and sensitization of bronchopulmonary C-fibers shown by increased firing rate. Upregulation of TRPV1 was associated with increased gene expression of the tropomyosin receptor kinase A (TrkA) in the nodose/jugular ganglia.

13.4. Nonneuronal TRP channels causing disease

As described in Section 13.2, TRP channels are expressed in several nonneuronal cells in the airway including epithelial cells, smooth muscle cells, fibroblasts and inflammatory cells. It is important to note that increasing evidence highlights the role of these TRP channels in airway disease. Increased airway smooth muscle contractility and subsequent airway narrowing is associated with obstructive disease states including asthma and COPD and has been shown to be due in part to increased activation of sensory afferents. Airway smooth muscle has also been shown to undergo structural changes contributing to airway modeling, and to orchestrate the inflammatory process in chronic airway disease. TRP channels expressed by ASM cells including members of the TRPC, TRPM, TRPV, and TRPA superfamilies are thought to be important for regulating abnormal SM structural functional changes during disease (Ong et al., 2003; Jha et al., 2015). Recent studies in the cystic fibrosis airway have shown that TRPC6-mediated Ca2+ influx was abnormally increased in cystic fibrosis (CF) epithelial cells (Antigny et al., 2011). Authors conclude that abnormal Ca2+ signaling contributing to CF airway pathophysiology is due to a loss of functional coupling between TRPC6 and CFTR, which is dysfunctional in CF. A recent study has identified a role for TRPA1 in the inflammatory response to infection in the CF airway. Results showed that inhibition of TRPA1 expression resulted in a reduction of release of IL-8, IL-1β, and TNF-α, from CF primary bronchial epithelial cells exposed to Pseudomonas aeruginosa (Prandini et al., 2016) TRPV4 channels have recently been implicated in airway pathologies including acute lung injury (ALI), idiopathic pulmonary fibrosis (IPF), and pulmonary hypertention (De Logu, 2016). TRPV4 expressed by neutrophils has been identified as a novel regulator of neutrophil activation and response to proinflammatory stimuli in ALI pathophysiology (Yin et al., 2016). The TRPV4 channel has also been linked to pulmonary fibrogenesis in IPF as the mechanical sensor that controls myofibroblast differentiation (Rahaman et al., 2014). TRPV4 is also thought to be involved in the development of pulmonary hypertension due to the role of this channel in vasoconstriction and regulation of arterial tone (Xia et al., 2013).

13.5. Conclusion

In summary, TRP channels play an integral part in protecting the airway from environmental challenges. It is unclear what causes the transition from early defensive immune and inflammatory responses to chronic responses and disease pathology, but it is thought that TRP channels may be at least in part responsible. Enhanced neuronal sensory responses in the diseased airway may be caused by increased expression of TRP channels or altered expression patterns resulting in a change in phenotype. Increased channel activity may be a result of sensitization, increasing excitability of afferent nerves reducing the threshold for activation or alternatively an increase in inflammatory mediators that can activate TRP channels. Further work will help to understand the communication between neurogenic and immune responses and the development of new therapeutic strategies.

References

  • Abdullah, H. et al. 2014. Rhinovirus upregulates transient receptor potential channels in a human neuronal cell line: Implications for respiratory virus-induced cough reflex sensitivity. Thorax, 69(1): 46–54. [PubMed: 24002057]
  • Adelroth, E. et al. 1986. Airway responsiveness to leukotrienes C4 and D4 and to methacholine in patients with asthma and normal controls. N Engl J Med, 315: 480–484. [PubMed: 3526153]
  • Alenmyr, L. et al. 2009. TRPV1-mediated itch in seasonal allergic rhinitis. Allergy, 64(5): 807–810. [PubMed: 19220220]
  • Alenmyr, L. et al. 2012. Effect of mucosal TRPV1 inhibition in allergic rhinitis. Basic Clin Pharmacol Toxicol, 110(3): 264–268. [PubMed: 21951314]
  • Alenmyr, L. et al. 2014. TRPV4-mediated calcium influx and ciliary activity in human native airway epithelial cells. Basic Clin Pharmacol Toxicol, 114: 210–216. [PubMed: 24034343]
  • Alvarez, D.F. et al. 2006. Transient receptor potential vanilloid 4-mediated disruption of the alveolar septal barrier: A novel mechanism of acute lung injury. Circ Res, 99: 988–995. [PMC free article: PMC2562953] [PubMed: 17008604]
  • Anand, P. and K. Bley. 2011. Topical capsaicin for pain management: Therapeutic potential and mechanisms of action of the new high-concentration capsaicin 8% patch. Br J Anaesth, 107: 490–502. [PMC free article: PMC3169333] [PubMed: 21852280]
  • Andersson, D.A. et al. 2008. Transient receptor potential A1 is a sensory receptor for multiple products of oxidative stress. J Neurosci, 28(10): 2485–2494. [PMC free article: PMC2709206] [PubMed: 18322093]
  • Andrè, E., B. Campi, and S. Materazzi. 2008. Cigarette smoke–induced neurogenic inflammation is mediated by α, β-unsaturated aldehydes and the TRPA1 receptor in rodents. J Clin Invest, 118: 2574–2582. [PMC free article: PMC2430498] [PubMed: 18568077]
  • Andrè, E. et al. 2009. Transient receptor potential ankyrin receptor 1 is a novel target for pro-tussive agents. Br J Pharmacol, 158: 1621–1628. [PMC free article: PMC2795228] [PubMed: 19845671]
  • Antigny, F. et al. 2011. Transient receptor potential canonical channel 6 links Ca2+ mishandling to cystic fibrosis transmembrane conductance regulator channel dysfunction in cystic fibrosis. Am J Respir Cell Mol Biol, 44(1): 83–90. [PubMed: 20203293]
  • Arden, K. E. et al. 2006. Frequent detection of human rhinoviruses, paramyxoviruses, coronaviruses, and bocavirus during acute respiratory tract infections. J Med Virol, 78: 1232–1240. [PMC free article: PMC7167201] [PubMed: 16847968]
  • Bandell, M. et al. 2004. Noxious cold ion channel TRPA1 is activated by pungent compounds and bradykinin. Neuron, 41: 849–857. [PubMed: 15046718]
  • Barnes, P.J. 2016. Inflammatory mechanisms in patients with chronic obstructive pulmonary disease. J Allergy Clin Immunol, 138(1):16-27. [PubMed: 27373322]
  • Bateman, E.D. et al. 2009. Alternative mechanisms for tiotropium. Pulm Pharmacol Ther, 22(6): 533–542. [PubMed: 19635581]
  • Bautista, D.M. et al. 2005. Pungent products from garlic activate the sensory ion channel TRPA1. Proc Natl Acad Sci USA, 102: 12248–12252. [PMC free article: PMC1189336] [PubMed: 16103371]
  • Bautista, D.M. et al. 2006. TRPA1 mediates the inflammatory actions of environmental irritants and proalgesic agents. Cell, 124(6): 1269–1282. [PubMed: 16564016]
  • Baxter, M. et al. 2014. Role of transient receptor potential and pannexin channels in cigarette smoke-triggered ATP release in the lung. Thorax, 69: 1080–1089. [PubMed: 25301060]
  • Bertin, S. et al. 2014. The ion channel TRPV1 regulates the activation and proinflammatory properties of CD4+ T cells. Nat Immunol, 15: 1055–1063. [PMC free article: PMC4843825] [PubMed: 25282159]
  • Bessac, B.F. et al. 2008. TRPA1 is a major oxidant sensor in murine airway sensory neurons. J Clin Invest, 118(5): 1899–1910. [PMC free article: PMC2289796] [PubMed: 18398506]
  • Bhattacharya, A. et al. 2007. Pharmacology and antitussive efficacy of 4-(3-trifluoromethyl-pyridin-2-yl)-peperazine-1-carboxylic acid (5-trifluoromethyl-pyridin-2-yl)-amide (JNJ17203212), a transient receptor potential vanilloid 1 antagonist in guinea pigs. J Pharmacol Exp Ther, 323: 665–674. [PubMed: 17690251]
  • Bhattacharya, M.R. et al. 2008. Radial stretch reveals distinct populations of mechanosensitive mammalian somatosensory neurons. Proc Natl Acad Sci USA, 105: 20015–20020. [PMC free article: PMC2604979] [PubMed: 19060212]
  • Birrell, M.A. et al. 2009. TRPA1 agonists evoke coughing in guinea-pig and human volunteers. Am J Respir Crit Care Med, 180: 1042–1047. [PMC free article: PMC2784411] [PubMed: 19729665]
  • Birrell, M.A. et al. 2014. Tiotropium modulates transient receptor potential V1 (TRPV1) in airway sensory nerves: A beneficial off-target effect? J Allergy Clin Immunol, 133(3): 679–687. [PMC free article: PMC3969581] [PubMed: 24506933]
  • BonvinI, S.J. et al 2016. Transient receptor potential cation channel, subfamily V, member 4 and airway sensory afferent activation: Role of adenosine triphosphate. J Allergy Clin Immunol, 138(1): 249–261. [PMC free article: PMC4929136] [PubMed: 26792207]
  • Brozmanova, M. et al. 2012. Comparison of TRPA1-versus TRPV1-mediated cough in guinea pigs. Eur J Pharmacol, 689: 211–218. [PMC free article: PMC4667741] [PubMed: 22683866]
  • Büch, T.R.H. et al. 2013. Functional expression of the transient receptor potential channel TRPA1, a sensor for toxic lung inhalants, in pulmonary epithelial cells. Chem Biol Interact, 206(3): 462–471. [PubMed: 23994502]
  • Buday, T. et al. 2012. Modulation of cough response by sensory inputs from the nose—Role of trigeminal TRPA1 versus TRPM8 channels. Cough, 8: 11. [PMC free article: PMC3546011] [PubMed: 23199233]
  • Caceres, A.I. et al. 2009. A sensory neuronal ion channel essential for airway inflammation and hyperreactivity in asthma. Proc Natl Acad Sci U S A, 106: 9099–9104. [PMC free article: PMC2684498] [PubMed: 19458046]
  • Canning, B.J. 2011. Functional implications of the multiple afferent pathways regulating cough. Pulm Pharmacol Ther, 24: 295–299. [PubMed: 21272660]
  • Canning, B.J., N. Mori, and S.B. Mazzone. 2006. Vagal afferent nerves regulating the cough reflex. Respir Physiol Neurobiol, 152(3): 223–242. [PubMed: 16740418]
  • Cantero-Recasens, G. et al. 2010. Loss of function of transient receptor potential vanilloid 1 (TRPV1) genetic variant is associated with lower risk of active childhood asthma. J Biol Chem, 285: 27532–27535. [PMC free article: PMC2934619] [PubMed: 20639579]
  • Cheah, E.Y. et al. 2014. Acrolein relaxes mouse isolated tracheal smooth muscle via a TRPA1-dependent mechanism. Biochem Pharmacol, 89(1): 148–156. [PubMed: 24561178]
  • Deering-Rice, C.E. et al. 2012. Transient receptor potential vanilloid-1 (TRPV1) is a mediator of lung toxicity for coal fly ash particulate material. Mol Pharmacol, 81: 411–419. [PMC free article: PMC3286291] [PubMed: 22155782]
  • Deering-Rice, C.E. et al. 2015. Activation of transient receptor potential ankyrin-1 by insoluble particulate material and association with asthma. Am J Respir Cell Mol Biol, 53(6): 893–901. [PMC free article: PMC4742944] [PubMed: 26039217]
  • Delany, N.S. et al. 2001. Identification and characterization of a novel human vanilloid receptor-like protein, VRL-2. Physiol Genomics, 4(3): 165–174. [PubMed: 11160995]
  • De Logu, F. et al. 2016 TRP functions in the broncho-pulmonary system. Semin Immunopathol, 38(3): 321–329. [PubMed: 27083925]
  • Delescluse, I., H. Mace, and J.J. Adcock. 2012. Inhibition of airway hyper-responsiveness by TRPV1 antagonists (SB-705498 andPF-04065463) in the unanaesthetized, ovalbumin-sensitized guinea pig. Br J Pharmacol, 166: 1822–1832. [PMC free article: PMC3402807] [PubMed: 22320181]
  • Devos, F.C. et al. 2016. Neuro-immune interactions in chemical-induced airway hyperreactivity TRPA1 channels–bronchoconstriction and AHR Eur Respir J, 48(2): 380–392. [PubMed: 27126687]
  • Dietrich, A. et al. 2006. Cation channels of the transient receptor potential superfamily: Their role in physiological and pathophysiological processes of smooth muscle cells. Pharmacol Ther, 112: 744–760. [PubMed: 16842858]
  • Doherty, M.J. et al. 2000.Capsaicin responsiveness and cough in asthma and chronic obstructive pulmonary disease. Thorax, 55(8): 643–649. [PMC free article: PMC1745828] [PubMed: 10899239]
  • Fernández-Fernández, J.M. et al. 2008. Functional coupling of TRPV4 cationic channel and large conductance, calcium-dependent potassium channel in human bronchial epithelial cell lines. Pflugers Arch, 457: 149–159. [PubMed: 18458941]
  • Garcia-Elias, A. et al. 2014. The TRPV4 channel. Handb Exp Pharmacol, 222: 293–319. [PubMed: 24756711]
  • Gauvreau, G.M., A.I. El-Gammal, and P.M. O'Byrne. 2015. Allergen-induced airway responses. Eur Respir J, 46(3): 819–831. [PubMed: 26206871]
  • Grace, M.S. and M.G. Belvisi. 2011. TRPA1 receptors in cough. Pulm Pharmacol Ther, 24(3): 286–288. [PubMed: 21074632]
  • Grace, M. et al. 2012. Transient receptor potential channels mediate the tussive response to prostaglandin E2 and bradykinin. Thorax, 67: 891–900. [PMC free article: PMC3446777] [PubMed: 22693178]
  • Grace, M.S. et al. 2014. Transient receptor potential (TRP) channels in the airway: Role in airway disease. Br J Pharmacol, 171(10): 2593–2607. [PMC free article: PMC4009002] [PubMed: 24286227]
  • Groneberg, D.A. et al. 2004. Increased expression of transient receptor potential vanilloid-1 in airway nerves of chronic cough. Am J Respir Crit Care Med, 170: 1276–1280. [PubMed: 15447941]
  • Hamanaka, K., M.Y. Jian, and M.I. Townsley. 2010. TRPV4 channels augment macrophage activation and ventilator-induced lung injury. Am J Physiol Lung Cell Mol Physiol, 299(3): L353–L362. [PMC free article: PMC2951075] [PubMed: 20562229]
  • Haque, R.A., O.S. Usmani, and P.J. Barnes. 2005. Chronic idiopathic cough: A discrete clinical entity? Chest, 127: 1710–1713. [PubMed: 15888850]
  • Hinman, A. et al. 2006. TRP channel activation by reversible covalent modification. Proc Natl Acad Sci U S A, 103: 19564–19568. [PMC free article: PMC1748265] [PubMed: 17164327]
  • Holgate, S.T. 2008. Pathogenesis of asthma. Clin Exp Allergy, 38(6): 872–897. [PubMed: 18498538]
  • Jang, Y. et al. 2012. Quantitative analysis of TRP channel genes in mouse organs. Arch Pharm Res, 35(10): 1823–1830. [PubMed: 23139135]
  • Jha, A. et al. 2015. A role for transient receptor potential ankyrin 1 cation channel (TRPA1) in airway hyper-responsiveness? Can J Physiol Pharmacol, 93(3): 171–176. [PubMed: 25654580]
  • Jia, Y. et al. 2004. Functional TRPV4 channels are expressed in human airway smooth muscle cells. Am J Physiol Lung Cell Mol Physiol, 287(2): L272–L278. [PubMed: 15075247]
  • Jordt, S.E. et al. 2004. Mustard oils and cannabinoids excite sensory nerve fibres through the TRP channel ANKTM1. Nature, 427: 260–265. [PubMed: 14712238]
  • Kenyon, N.J. et al. 2012. Occupational asthma. Clin Rev Allergy Immunol, 43(1–2): 3–13. [PubMed: 21573916]
  • Laumbach, R.J. and H.M. Kipen. 2012. Respiratory health effects of air pollution: Update on biomass smoke and traffic pollution. J Allergy Clin Immunol, 129: 3–11. [PMC free article: PMC3272333] [PubMed: 22196520]
  • Lee, L.Y. et al. 2015. Interaction between TRPA1 and TRPV1: Synergy on pulmonary sensory nerves. Pulm Pharmacol Ther, 35: 87–93. [PMC free article: PMC4690745] [PubMed: 26283426]
  • Lieu, T.M. et al. 2012. TRPV1 induction in airway vagal low-threshold mechanosensory neurons by allergen challenge and neurotrophic factors. Am J Physiol Lung Cell Mol Physiol, 302: L941–L948. [PMC free article: PMC3362153] [PubMed: 22345578]
  • Li, J. et al. 2011. TRPV4-mediated calcium influx into human bronchial epithelia upon exposure to diesel exhaust particles. Environ Health Perspect, 119(6): 784–793. [PMC free article: PMC3114812] [PubMed: 21245013]
  • Lin, Y.S. et al. 2010. Activations of TRPA1 and P2X receptors are important in ROS-mediated stimulation of capsaicin-sensitive lung vagal afferents by cigarette smoke in rats. J Appl Physiol, 108: 1293–1303. [PubMed: 20167675]
  • Lorenzo, I.M. et al. 2008. TRPV4 channel participates in receptor-operated calcium entry and ciliary beat frequency regulation in mouse airway epithelial cells. Proc Natl Acad Sci U S A, 105(34): 12611–12616. [PMC free article: PMC2527959] [PubMed: 18719094]
  • Macpherson, L.J. et al. 2007. Noxious compounds activate TRPA1 ion channels through covalent modification of cysteines. Nature, 445: 541–545. [PubMed: 17237762]
  • McAlexander, M.A. et al. 2014. Transient receptor potential vanilloid 4 activation constricts the human bronchus via the release of cysteinyl leukotrienes. J Pharmacol Exp Ther, 349(1): 118–125. [PMC free article: PMC3965887] [PubMed: 24504097]
  • McGarvey, L.P. et al. 2014. Increased expression of bronchial epithelial transient receptor potential vanilloid 1 channels in patients with severe asthma. J Allergy Clin Immunol, 133(3): 704–712. [PubMed: 24210884]
  • Mitchell, J.E. et al. 2005. Expression and characterization of the intracellular vanilloid receptor (TRPV1) in bronchi from patients with chronic cough. Exp Lung Res, 31: 295–306. [PubMed: 15962710]
  • Mori, T. 2011. Lack of transient receptor potential vanilloid-1 enhances Th2-biased immune response of the airways in mice receiving intranasal, but not intraperitoneal, sensitization. Int Arch Allergy Immunol, 156: 305–312. [PMC free article: PMC3130892] [PubMed: 21720176]
  • Mukhopadhyay, P. et al. 2011. Expression of functional TRPA1 receptor on human lung fibroblast and epithelial cells. J Recept Signal Transduct Res, 31(5): 350–358. [PubMed: 21848366]
  • Naeher, L.P. et al. 2007. Woodsmoke health effects. Inhal Toxicol, 19: 67–106. [PubMed: 17127644]
  • Nakajima, T. et al. 2006. Cough sensitivity in pure cough variant asthma elicited using continuous capsaicin inhalation. Allergol Int, 55: 149–155. [PubMed: 17075251]
  • Nassenstein, C. et al. 2008. Expression and function of the ion channel TRPA1 in vagal afferent nerves innervating mouse lungs. J Physiol, 586: 1595–1604. [PMC free article: PMC2375701] [PubMed: 18218683]
  • Nassini, R. et al. 2012. Transient receptor potential ankyrin 1 channel localized to non-neuronal airway cells promotes non-neurogenic inflammation. PLOS ONE, 7(8): e42454. [PMC free article: PMC3419223] [PubMed: 22905134]
  • Nicholson, K.G., J. Kent, and D.C. Ireland. 1993. Respiratory viruses and exacerbations of asthma in adults. BMJ, 307: 982–986. [PMC free article: PMC1679193] [PubMed: 8241910]
  • O'Byrne, P.M., G.M. Gauvreau, and J.D. Brannan. 2009. Provoked models of asthma: What have we learnt? Clin Exp Allergy, 39: 181e92. [PubMed: 19187330]
  • Ong, H.L. et al. 2003. Evidence for the expression of transient receptor potential proteins in guinea pig airway smooth muscle cells. Respirology, 8(1): 23–32. [PubMed: 12856738]
  • Parenti, A. et al. 2016. What is the evidence for the role of TRP channels in inflammatory and immune cells? Br J Pharmacol, 173(6): 953–969. [PMC free article: PMC5341240] [PubMed: 26603538]
  • Pauwels, R.A. et al. 2001. Global strategy for the diagnosis, management, and prevention of chronic obstructive pulmonary disease. NHLBI/WHO global initiative for chronic obstructive lung disease (GOLD) workshop summary. Am J Respir Crit Care Med, 163: 1256–1276. [PubMed: 11316667]
  • Pecova, R. et al. 2008. Cough reflex sensitivity testing in seasonal allergic rhinitis patients and healthy volunteers. J Physiol Pharmacol, 59(6): 557–564. [PubMed: 19218681]
  • Plevkova, J. et al. 2013. The role of trigeminal nasal TRPM8-expressing afferent neurons in the antitussive effects of menthol. J Appl Physiol, 115: 268–274. [PMC free article: PMC3727002] [PubMed: 23640596]
  • Prandini, P. et al. 2016. TRPA1 channels modulate inflammatory response in respiratory cells from cystic fibrosis patients. Am J Respir Cell Mol Biol, 55(5): 645–656. [PubMed: 27281024]
  • Preti, D., A. Szallasi, and R. Patacchini. 2012. TRP channels as therapeutic targets in airway disorders: A patent review. Expert Opin Ther Pat, 22: 663–695. [PubMed: 22667456]
  • Raemdonck, K. et al. 2012. A role for sensory nerves in the late asthmatic response. Thorax, 67: 19–25. [PubMed: 21841185]
  • Rahaman, S.O. et al. 2014. TRPV4 mediates myofibroblast differentiation and pulmonary fibrosis in mice. J Clin Invest, 124: 5225–5238. [PMC free article: PMC4348970] [PubMed: 25365224]
  • Rehman, R. et al. 2013 TRPV1 inhibition attenuates IL-13 mediated asthma features in mice by reducing airway epithelial injury. Int Immunopharmacol, 15: 597–605. [PubMed: 23453702]
  • Ruan, T. et al. 2014. Sensitization by pulmonary reactive oxygen species of rat vagal lung C-fibers: The roles of the TRPV1, TRPA1, and P2X receptors. PLOS ONE, 9: e91763. [PMC free article: PMC3974698] [PubMed: 24699274]
  • Saetta, M. et al. 1997. Inflammatory cells in the bronchial glands of smokers with chronic bronchitis. Am J Respir Crit Care Med, 156: 1633–1639. [PubMed: 9372687]
  • Saetta, M. et al. 1998. CD8+ T-lymphocytes in peripheral airways of smokers with chronic obstructive pulmonary disease. Am J Respir Crit Care Med, 157(3): 822–826. [PubMed: 9517597]
  • Sabnis, A.S. et al. 2008. Human lung epithelial cells express a functional cold-sensing TRPM8 variant. Am J Respir Cell Mol Biol, 39(4): 466–474. [PMC free article: PMC2551706] [PubMed: 18458237]
  • Salvi, S.S. and P.J. Barnes. 2009. Chronic obstructive pulmonary disease in non-smokers. Lancet, 374: 733–743. [PubMed: 19716966]
  • Samivel, R. et al. 2016. The role of TRPV1 in the CD4+ T cell mediated inflammatory response of allergic rhinitis. Oncotarget, 7(1): 14860. [PMC free article: PMC4807989] [PubMed: 26700618]
  • Satoh, J. and M. Yamakage. 2009. Desflurane induces airway contraction mainly by activating transient receptor potential A1 of sensory C-fibers. J Anesth, 23(4): 620–623. [PubMed: 19921381]
  • Selvarajah, S. et al. 2016. Multiple circulating cytokines are coelevated in chronic obstructive pulmonary disease. Mediators Inflamm, 2016:3604842. [PMC free article: PMC4976159] [PubMed: 27524865]
  • Shapiro, D. et al. 2013. Activation of transient receptor potential ankyrin-1 (TRPA1) in lung cells by wood smoke particulate material. Chem Res Toxicol, 26: 750–758. [PMC free article: PMC3670828] [PubMed: 23541125]
  • Smit L.A. et al. 2012. Transient receptor potential genes, smoking, occupational exposures and cough in adults. Respir Res, 13: 26. [PMC free article: PMC3342106] [PubMed: 22443337]
  • Spina, D. and C.P. Page. 2013. Regulating cough through modulation of sensory nerve function in the airways. Pulm Pharmacol Ther, 26(5): 486–490. [PubMed: 23524012]
  • Story, G.M. et al. 2003. ANKTM1, a TRP-like channel expressed in nociceptive neurons, is activated by cold temperatures. Cell, 112(6): 819–829. [PubMed: 12654248]
  • Talavera, K. et al. 2009. Nicotine activates the chemosensory cation channel TRPA1. Nat Neurosci, 12: 1293–1300. [PubMed: 19749751]
  • Trevisani, M. et al. 2007. 4-Hydroxynonenal, an endogenous aldehyde, causes pain and neurogenic inflammation through activation of the irritant receptor TRPA1. Proc Natl Acad Sci U S A, 104: 13519–13524. [PMC free article: PMC1948902] [PubMed: 17684094]
  • Van Gerven, L., G. Boeckxstaens, and P. Hellings. 2012. Up-date on neuro-immune mechanisms involved in allergic and non-allergic rhinitis. Rhinology, 50: 227–235. [PubMed: 22888478]
  • Van Gerven, L. et al. 2014. Capsaicin treatment reduces nasal hyperreactivity and transient receptor potential cation channel subfamily V, receptor 1 (TRPV1) overexpression in patients with idiopathic rhinitis. J Allergy Clin Immunol, 133(5): 1332–1339. [PubMed: 24139494]
  • Vergnolle, N. et al. 2010. A role for transient receptor potential vanilloid 4 in tonicity-induced neurogenic inflammation. Br J Pharmacol, 159(5): 1161–1173. [PMC free article: PMC2839274] [PubMed: 20136846]
  • Watanabe, H., J.B. Davis, and D. Smart. 2002a. Activation of TRPV4 channels (hVRL-2/mTRP12) by phorbol derivatives. J Biol Chem, 277: 13569–13577. [PubMed: 11827975]
  • Watanabe, N. et al. 2006. Immunohistochemical co-localization of transient receptor potential vanilloid (TRPV) 1 and sensory neuropeptides in the guinea-pig respiratory system. Neuroscience, 141: 1533–1543. [PubMed: 16765524]
  • Widdicombe, J.G. 1995. Neurophysiology of the cough reflex. Eur Respir J, 8: 1193–1202. [PubMed: 7589405]
  • Xia, Y. et al. 2013. TRPV4 channel contributes to serotonin-induced pulmonary vasoconstriction and the enhanced vascular reactivity in chronic hypoxic pulmonary hypertension. Am J Physiol Cell Physiol, 305: C704–C715. [PMC free article: PMC3798671] [PubMed: 23739180]
  • Xing, H. et al. 2008. TRPM8 mechanism of autonomic nerve response to cold in respiratory airway. Mol Pain, 5(4): 22. [PMC free article: PMC2430548] [PubMed: 18534015]
  • Yin, J. et al. 2016. Role of transient receptor potential vanilloid 4 in neutrophil activation and acute lung injury. Am J Respir Cell Mol Biol, 54(3): 370–383. [PubMed: 26222277]
  • Zhang, G. et al. 2008. Altered expression of TRPV1 and sensitivity to capsaicin in pulmonary myelinated afferents following chronic airway inflammation in the rat. J Physiol, 586: 5771–5786. [PMC free article: PMC2655410] [PubMed: 18832423]
  • Zhang, L. et al. 2016. Activation of cold-sensitive channels TRPM8 and TRPA1 inhibits the proliferative airway smooth muscle cell phenotype. Lung, 194(4): 595–603. [PubMed: 27236325]
  • Zhao, L. et al. 2013. Effect of TRPV1 channel on the proliferation and apoptosis in asthmatic rat airway smooth muscle cells. Exp Lung Res, 39(7): 283–294. [PubMed: 23919305]
  • Zheng, X. et al 2013. Arachidonic acid-induced dilation in human coronary arterioles: Convergence of signaling mechanisms on endothelial TRPV4- mediated Ca2+ entry. J Am Heart Assoc, 2: e000080. [PMC free article: PMC3698766] [PubMed: 23619744]
  • Zhuang, J., L. Zhao, and F. Xu. 2014. Maternal nicotinic exposure produces a depressed hypoxic ventilatory response and subsequent death in postnatal rats. Physiol Rep, 2: 1–12. [PMC free article: PMC4098749] [PubMed: 24872357]
  • Zhuang, J. et al. 2015. Prenatal nicotinic exposure augments cardiorespiratory responses to activation of bronchopulmonary C-fibers. Am J Physiol Lung Cell Mol Physiol, 308: L922–L930. [PMC free article: PMC4421788] [PubMed: 25747962]
  • Zhu, G. et al. 2009. Association of TRPV4 gene polymorphisms with chronic obstructive pulmonary disease. Hum Mol Genet, 18(11): 2053–2062. [PubMed: 19279160]
  • Zurgorg, S. et al. 2007. Direct activation of the ion channel TRPA1 by Ca2þ. Nat Neurosci, 10: 277e9. [PubMed: 17259981]
© 2018 by Taylor & Francis Group, LLC.
Bookshelf ID: NBK476121PMID: 29356492DOI: 10.4324/9781315152837-13

Views

  • PubReader
  • Print View
  • Cite this Page

Other titles in this collection

Related information

  • PMC
    PubMed Central citations
  • PubMed
    Links to PubMed

Similar articles in PubMed

See reviews...See all...

Recent Activity

Your browsing activity is empty.

Activity recording is turned off.

Turn recording back on

See more...