Skip to main content
The Journal of Physiology logoLink to The Journal of Physiology
. 2010 Dec 6;589(Pt 7):1515–1525. doi: 10.1113/jphysiol.2010.201855

TRPM2: a multifunctional ion channel for calcium signalling

Adriana Sumoza-Toledo 1,2, Reinhold Penner 1,3,4
PMCID: PMC3099011  PMID: 21135052

Abstract

The transient potential receptor melastatin-2 (TRPM2) channel has emerged as an important Ca2+ signalling mechanism in a variety of cells, contributing to cellular functions that include cytokine production, insulin release, cell motility and cell death. Its ability to respond to reactive oxygen species has made TRPM2 a potential therapeutic target for chronic inflammation, neurodegenerative diseases, and oxidative stress-related pathologies. TRPM2 is a non-selective, calcium (Ca2+)-permeable cation channel of the melastatin-related transient receptor potential (TRPM) ion channel subfamily. It is activated by intracellular adenosine diphosphate ribose (ADPR) through a diphosphoribose hydrolase domain in its C-terminus and regulated through a variety of factors, including synergistic facilitation by [Ca2+]i, cyclic ADPR, H2O2, NAADP, and negative feedback regulation by AMP and permeating protons (pH). In addition to its role mediating Ca2+ influx into the cells, TRPM2 can also function as a lysosomal Ca2+ release channel, contributing to cell death. The physiological and pathophysiological context of ROS-mediated events makes TRPM2 a promising target for the development of therapeutic tools of inflammatory and degenerative diseases.


Adriana Sumoza-Toledo, PhD (left) is a Scientist at the Queen's Medical Center and adjunct Assistant Professor at the University of Hawaii in Honolulu, HI. She earned her Ph.D. in Molecular Biomedicine from CINVESTAV, Mexico and received post-doctoral training in Immunology and Biophysics at Nationwide Children's Hospital, Columbus, OH and The Queen's Medical Center, Honolulu, HI, respectively. Her research background includes cell cytoskeleton, cell signaling, immunology, biochemistry, and electrophysiology. Her research interests focus on TRPM2 channels in innate immune cell function and cancer biology. Reinhold Penner, MD, PhD (right) is the Director of Research at The Center for Biomedical Research at The Queen's Medical Center in Honolulu, Hawaii. He worked at the Max-Planck-Institute for Biophysical Chemistry in Göttingen, Germany for 12 years and relocated to Honolulu in 1997 to establish the Laboratory of Cell and Molecular Signaling at The Queen's Medical Center. His work has contributed to the understanding of store-operated calcium entry and led to the discovery the Calcium Release-Activated Calcium (CRAC) channels. His laboratory also contributed to the discovery and characterization of several TRP channels, including TRPM2, TRPM4, TRPM5, and TRPM7. His group is now involved in drug discovery efforts for a range of ion channels involved in inflammation, autoimmune diseases, diabetes, and cancer.

graphic file with name tjp0589-1515-m1.jpg

Introduction

Transient receptor potential (TRP) proteins represent a large superfamily of six-transmembrane (6TM) monovalent and divalent cation-permeable ion channels that are homologues of the Drosophila melanogaster TRP protein, a Ca2+-permeable channel that is essential for phototransduction (Ramsey et al. 2006; Nilius, 2007; Venkatachalam & Montell, 2007). The history of TRP proteins began with the discovery of a mutant strain of Drosophila melanogaster that had an abnormal response to prolonged illumination and therefore was visually impaired (Cosens & Manning, 1969). This led to the isolation of the first TRP protein in 1989 (Montell & Rubin, 1989) and subsequent identification of mammalian TRP-like proteins (Ramsey et al. 2006; Nilius, 2007; Venkatachalam & Montell, 2007).

Currently, the mammalian TRP superfamily is composed of 28 members which are grouped into six subfamilies based on their amino acid sequence homology: canonical or classic (C), vanilloid (V), melastatin (M), polycystin (P), mucolipin (ML) and ankyrin (A). Typically, the TRP protein structure is characterized by six-transmembrane domains, a pore region situated between the fifth and sixth transmembrane domains, with N- and C-termini oriented toward the cytoplasm (Ramsey et al. 2006; Nilius, 2007; Venkatachalam & Montell, 2007; Gaudet, 2008, 2009; Latorre et al. 2009; Song & Yuan, 2010). To form a channel, TRP proteins typically assemble into homo-tetramers, although in some cases hetero-tetramers have been demonstrated to form functional channels (Goel et al. 2002; Hofmann et al. 2002; Li et al. 2006).

Even within the major subfamilies, TRP channels exhibit a remarkable diversity in their mechanisms of activation (thermally activated, receptor activated and ligand activated), selectivity and permeability to cations (Na+, Ca2+, Mg2+, K+) (Ramsey et al. 2006; Nilius, 2007; Venkatachalam & Montell, 2007). Activation of TRP channels in the plasma membrane mediates the flux of Ca2+ and/or Na+ into the cell, thereby raising intracellular concentration of these ions and depolarizing the cell (Ramsey et al. 2006; Nilius, 2007; Venkatachalam & Montell, 2007). Additionally, intracellular TRP channels may function as Ca2+ release channels and modulate the intracellular Ca2+ concentration ([Ca2+]i) (Lange et al. 2009; Dong et al. 2010; Gees et al. 2010). TRP channels are widely expressed in different tissues and cell types from worms, fruit flies, zebrafish, mice and humans (Venkatachalam & Montell, 2007), including excitable and non-excitable cells, and they can be found in the plasma membrane and organelles, demonstrating the diversity of their biological roles (Ramsey et al. 2006; Nilius, 2007; Venkatachalam & Montell, 2007). They are involved in sensory functions such as taste transduction and temperature sensation (Talavera et al. 2008), in homeostatic functions like Ca2+ and Mg2+ reabsorption and osmoregulation, or cellular functions like cell motility and muscle contraction (Venkatachalam & Montell, 2007; Gees et al. 2010).

The TRP-melastatin subfamily (TRPM), named for the tumour suppressor melastatin (TRPM1), contains eight mammalian members: TRPM1–TRPM8, which are divided into four homologue pairs: TRPM1/TRPM3, TRPM2/TRPM8, TRPM4/TRPM5 and TRPM6/TRPM7 (Fleig & Penner, 2004a,b;). The N-terminus of the TRPM subfamily members is characterized by four stretches of residues, designated as the TRPM homology domain or MHD (Perraud et al. 2001). Additionally, TRPM proteins present a TRP domain within their C-termini (LPPPFI) near the transmembrane segments (Venkatachalam & Montell, 2007). Several TRPM channels (TRPM1–6, and TRPM8) are expressed as multiple variants or isoforms, either as a full-length protein or a short variant, contributing to their biological complexity (Vazquez & Valverde, 2006; Venkatachalam & Montell, 2007). It has been suggested that these splice variants may play a role modulating the function, ion selectivity and localization of these channels. However, the significance of most of them has not yet been determined. In this review we focus on the TRPM2 channel and discuss its molecular and biophysical properties, as well as the physiological context in which it functions.

TRPM2 structure

TRPM2 (previously known as LTRPC2 or TRPC7) is a multifunctional Ca2+ permeable, non-selective cation channel with a unique C-terminal adenosine diphosphate ribose (ADPR) pyrophosphatase domain (Nudix-like domain or NUDT9 homology domain) (Perraud et al. 2001; Sano et al. 2001; Fleig & Penner, 2004a,b;). Like TRPM6/7, TRPM2 is known as a ‘chanzyme’ because of its dual function of ion channel and C-terminal enzyme domain. The human TRPM2 gene is located in chromosome 21q22.3, consists of 32 exons and spans approximately 90 kb (Nagamine et al. 1998). An additional 5′-exon contained within a CpG island has been identified in the human TRPM2 gene (Uemura et al. 2005). On the other hand, the mouse TRPM2 gene contains 34 exons and spans about 61 kb (Uemura et al. 2005). The human TRPM2 transcript is ∼6.5 kb and encodes a protein of 1503 amino acids (1507 in mouse and rat) with a predicted molecular mass of ∼170 kDa (Nagamine et al. 1998; Hara et al. 2002; Jiang et al. 2010) (Fig. 1). The TRPM2 protein structure comprises six transmembrane segments (S1–S6), flanked by the intracellular N- and C-termini, with the pore-forming loop domain located between S5 and S6 (Nagamine et al. 1998; Perraud et al. 2001; Sano et al. 2001). Additionally, the TRPM2 N-terminus has four homologous domains and a calmodulin (CaM) binding IQ-like motif, which plays a role in modulating channel activation (Perraud et al. 2001; Sano et al. 2001; McHugh et al. 2003; Fleig & Penner, 2004a,b; Tong et al. 2006) (Fig. 1). The significance of the MHD domains in TRPM2 function or expression remains to be established. On the other hand, the C-terminus contains a TRP box and a coil–coil domain (Jiang, 2007), which has been suggested to be critical for the homo-tetrameric assembly of TRPM2 (Fig. 1).

Figure 1. TRPM2 protein structure and transmembrane topology.

Figure 1

A, TRPM2 protein structure. Human TRPM2 is a protein of ∼170 kDa composed of 1503 amino acids (1507 in mouse and rat). The channel's N-terminal has four homologous regions (MHR) of unknown function and a calmodulin (CaM) binding IQ-like motif, followed by six transmembrane segments (TM: S1–S6). The TRPM2 pore-forming loop domain locates between S5 and S6. The TRPM2 C-terminus contains a TRP box and a coil–coil domain (CC), and a C-terminal adenosine diphosphate ribose (ADPR) pyrophosphatase domain (Nudix-like domain or NUDT9 homology domain, NUDT9-H). B, TRPM2 transmembrane topology. The TRPM2 N- and C-termini face the cytosol. Cytosolic ADPR binds to the TRPM2 NUDT9-H region and gates the channel, allowing calcium (Ca2+) and sodium (Na+) influx. ADPR is hydrolysed to ribose 5-phosphate and adenosine monophosphate (AMP) by TRPM2 NUDT9-H enzymatic activity. TRPM2 gating by ADPR is facilitated by hydrogen (H2O2), cyclic ADPR (cADPR) and Ca2+. AMP acts as a negative regulator of TRPM2 gating by ADPR and 8Br-cADPR inhibits cADPR- and H2O2-mediated effects.

ADPR is considered the primary gating molecule of TRPM2 (Perraud et al. 2001; Sano et al. 2001; Fleig & Penner, 2004a, b). It binds to the Nudix-like domain in the C-terminus of TRPM2 with high specificity and is subsequently hydrolysed (Km 100 μm, Vmax 100 nmol μg−1 min−1) to ribose 5-phosphate and AMP (Bessman et al. 1996; Perraud et al. 2001, 2003a,b) (Fig. 1). Although the physiological role of TRPM2 enzymatic activity has not been explored in detail, it is generally assumed that the enzymatic activity may serve to provide negative feedback inhibition for TRPM2 activity, since AMP antagonizes ADPR-mediated gating of TRPM2 (Kolisek et al. 2005; Beck et al. 2006; Lange et al. 2008). Upon binding of ADPR, TRPM2 channels will open and allow the permeation of sodium (Na+), potassium (K+) and Ca2+ into the cell with a relative permeability of PCa:PNa∼0.3–0.9 (Perraud et al. 2001; Sano et al. 2001; Venkatachalam & Montell, 2007) (Fig. 1). TRPM2 currents are characterized by a linear current–voltage (I–V) relationship with a reversal potential of ∼0 mV (Perraud et al. 2001; Sano et al. 2001). The single channel conductance is ∼60 pS with unusually long open times in the range of several seconds (Perraud et al. 2001).

TRPM2 gating by ADPR is subject to remarkable modulatory mechanisms, both inhibitory and facilitatory. Negative regulators include AMP (Kolisek et al. 2005; Beck et al. 2006; Lange et al. 2008) and protons (Du et al. 2009b; Starkus et al. 2010; Yang et al. 2010), whereas facilitation is observed with Ca2+ (McHugh et al. 2003; Starkus et al. 2007; Csanady & Torocsik, 2009), hydrogen peroxide (H2O2) (Hara et al. 2002; Kolisek et al. 2005), cyclic ADPR (cADPR) (Kolisek et al. 2005; Lange et al. 2008), and nicotinic acid adenine dinucleotide phosphate (NAADP) (Beck et al. 2006; Lange et al. 2008). Some of these modulatory effects appear to be directly mediated at the channel protein, whereas others are indirect, involving cytosolic components yet to be identified (Toth & Csanady, 2010).

TRPM2 is expressed at its highest in the brain but is also detected in other tissues such as bone marrow, spleen, heart, liver and lung, and in different cell types like pancreatic β-cells (Ishii et al. 2006b; Togashi et al. 2006; Lange et al. 2009; Uchida et al. 2010), endothelial cells (Hecquet et al. 2008; Hecquet & Malik, 2009; Hecquet et al. 2010), microglia (Kraft et al. 2004), neurons (Hill et al. 2006; Olah et al. 2009), cardiomyocytes (Yang et al. 2006), and immune cells (neutrophils, megakaryocytes, monocytes/macrophages) (Heiner et al. 2003; Carter et al. 2006; Lange et al. 2008; Yamamoto et al. 2008).

Although it was originally described as a plasma membrane channel, TRPM2 has recently been found to also function as a lysosomal Ca2+ release channel in pancreatic β-cells (Lange et al. 2009). Intracellular localization, albeit not in lysosomes, has also been reported for other TRP channels, including TRPV1 (Morenilla-Palao et al. 2004), TRPC5 (Bezzerides et al. 2004), TRPC3 (Singh et al. 2004), TRPM8 (Thebault et al. 2005), TRPML1–3 (Piper & Luzio, 2004; Cheng et al. 2010; Dong et al. 2010), and TRPM7 (Oancea et al. 2006). The factors that would determine the cellular localization TRPM2 and various other TRP channels remain to be defined, as well as whether the cellular localization serves a particular cellular function.

TRPM2 activation and modulation

ADPR is the most efficient TRPM2 activator, with variable EC50 values of 1–90 μm depending on the cell type investigated (Perraud et al. 2001; Sano et al. 2001; Inamura et al. 2003; Beck et al. 2006; Gasser et al. 2006; Starkus et al. 2007; Lange et al. 2008). The variability in EC50 values may arise from the modulatory mechanisms expressed in a given cell type. The main cellular pathways that generate free ADPR are the hydrolysis of NAD+ and/or cADPR by glycohydrolases, including the ectoenzymes CD38 and CD157 as well as the mitochondrial NADase (Lund et al. 1995, 1998; Lund, 2006; Malavasi et al. 2006). A further source of ADPR is provided by the combined action of poly(ADPR) polymerases (PARPs, PARP enzymes) and poly(ADPR) glycohydrolases (PARG enzymes), which indirectly generate ADPR via formation and hydrolysis of poly-ADPR when hyperactivated in response to DNA damage (Caiafa et al. 2009; Esposito & Cuzzocrea, 2009; Fauzee et al. 2010).

Other nucleotides with the ability to activate TRPM2 channels, although less effectively, are cADPR (EC50= 0.7 mm) (Kolisek et al. 2005; Lange et al. 2008), and NAADP (EC50= 0.73 mm) (Beck et al. 2006; Lange et al. 2008). Even though activation of TRPM2 by high concentrations of NAD+ (EC50= 1–1.8 mm) has been observed (Sano et al. 2001; Hara et al. 2002; Naziroglu & Luckhoff, 2008), its status as a direct agonist for TRPM2 remains to be established more thoroughly, since at least in some studies, contaminations with ADPR or metabolism of NAD+ may account for the observed TRPM2 activation by high concentrations of NAD+ (Beck et al. 2006; Grubisha et al. 2006). The concentrations of cADPR and NAADP required to activate TRPM2 are high compared with the physiological concentration of these nucleotides; however, these nucleotides can synergize with ADPR and increase TRPM2 sensitivity at much lower doses. Whether they bind directly to the Nudix domain or to different cooperative sites, or are converted to ADPR is not clearly understood. Nevertheless, abundant evidence suggests that TRPM2 activation is linked to pathways that involve the generation of these nucleotides (Fig. 2).

Figure 2. Signalling mechanisms for TRPM2 activation.

Figure 2

NAD+ and reactive oxygen species (ROS), including H2O2, accumulate during inflammation and tissue damage. External NAD+ may be converted to ADPR, cADPR and NAADP by the ectoenzymes CD38 and CD157. Extracellular ADPR may then bind to plasma membrane receptors (e.g. P2Y receptors) and increases [Ca2+]i through Ca2+ release from stores via G-proteins and phospholipase C (PLC) activation with subsequent IP3 production. H2O2 may also cross the plasma membrane and mobilize ADPR from mitochondria (both H2O2 and cADPR can synergize with ADPR to activate TRPM2). ADPR is also generated from poly-ADPR during ROS-induced DNA damage through activation of the PARP-1/PARG pathway. Free cytosolic ADPR will act on the NUDT9-H of lysosomal and plasma membrane TRPM2 channels, enabling Ca2+ influx across the plasma membrane and/or release of lysosomal Ca2+, raising the Ca2+ concentration in the cytosol. Ca2+ overload can trigger programmed cell death (apoptosis) and possibly necrosis. Finally, extracellular signals that remain to be identified could potentially induce the production of intracellular free ADPR, which may then gate TRPM2 channels in the lysosome and/or plasma membrane and regulate receptor-mediated signalling.

CD38, a multifunctional ectoenzyme widely expressed in haematopoietic cells and non-haematopoietic tissues, uses NAD+ as a substrate to catalyse the production of ADPR, cADPR and NAADP (Lund et al. 1995, 1998). CD38 knock-out (KO) neutrophils stimulated with the bacterial peptide formyl-methionyl-leucyl-phenylalanine (fMLP) present a reduced Ca2+ response when compared to wild-type cells (Partida-Sanchez et al. 2003). Similarly, fMLP-treated TRPM2 KO neutrophils have defects in Ca2+ influx (Yamamoto et al. 2008). Additionally, the fMLP-induced Ca2+ entry can be inhibited when neutrophils are treated with 8Br-ADPR or 8Br-cADPR (Partida-Sanchez et al. 2004; Partida-Sanchez et al. 2007). Interestingly, TRPM2 channels are expressed in the plasma membrane of neutrophils, suggesting a signalling pathway involving CD38–ADPR–TRPM2. Although ADPR is the main product of CD38 and evidence points to TRPM2 as the mediator of the Ca2+ entry, there are still open questions such as how the extracellular ADPR generated by CD38 crosses the plasma membrane and acts on the cytosolic Nudix domain of TRPM2 channels.

A novel acetyl-ADP ribose product, O-acetyl-ADP ribose (OAADPR), which is derived from the histone/protein deacetylase reaction mediated by sirtuins, induces TRPM2 currents by direct binding to the Nudix domain with a Kd of ∼100 μm (Grubisha et al. 2006; Tong & Denu, 2010). SIRT2 and SIRT3, mammalian sirtuins, have been suggested to generate the OAADPR that leads to TRPM2-dependent cell death induced by puromycin, since specific RNAi knockdown in TRPM2-expressing cells protects these cells from cell death (Grubisha et al. 2006). This suggests that, OAADPR may also function as a physiological regulator of TRPM2.

Extracellular Ca2+ and intracellular Ca2+ have also a critical role in the full activation of TRPM2 channels (McHugh et al. 2003; Starkus et al. 2007; Csanady & Torocsik, 2009). Intracellular Ca2+ facilitates TRPM2 activation by enhancing the channel sensitivity to ADPR. ADPR alone in the complete absence of Ca2+ cannot induce cation currents and a minimum of 30 nm internal Ca2+ is required to enable TRPM2 currents. However, this facilitation is not mimicked by Mg2+, Ba2+, or Zn2+ (Starkus et al. 2007). On the other hand, 200 μm external Ca2+ is as efficient as 1 mm Ca2+ in TRPM2 activation (Starkus et al. 2007). It has also been suggested that Ca2+ can gate the channel directly in a dose-dependent manner, with an EC50 of 17 μm, or 0.5 μm in the presence of 10 μm ADPR (McHugh et al. 2003; Starkus et al. 2007; Du et al. 2009a). Although the mechanism is still not clear it is thought to result from conformational changes due to [Ca2+]i-dependent tethering of CaM with TRPM2 IQ-like motif or other intracellular sites (Du et al. 2009a). However, other investigations did not observe Ca2+-induced activation (McHugh et al. 2003; Starkus et al. 2007; Csanady & Torocsik, 2009) and it is possible that Ca2+ at high concentrations could lead to ADPR production or ADPR release from mitochondria.

TRPM2 channels can also be activated by H2O2 (Hara et al. 2002; Kolisek et al. 2005) and other agents that produce oxygen and nitrogen species, although the mechanism of action remains unclear. It appears that the gating mechanism relies primarily on its ability to release ADPR from mitochondria (Ayub & Hallett, 2004). In addition, it has been shown that H2O2-mediated TRPM2 currents can be suppressed by reducing the ADPR concentration within the mitochondria (Perraud et al. 2005). Contrasting this hypothesis, a TRPM2 variant lacking the Nudix domain still responds to H2O2 (Zhang et al. 2003). Another source of ADPR under oxidative stress is the nucleus. ADPR generation in the nucleus involves the activation of PARP-1/PARG pathway by DNA damage. PARP-1 binds to damaged-DNA and catalyses the cleavage of NAD+ to nicotinamide and ADPR (Caiafa et al. 2009; Esposito & Cuzzocrea, 2009; Fauzee et al. 2010). ADPR is then polymerized onto various nuclear proteins, activating DNA repair mechanisms. Free ADPR is generated following the degradation of ADPR polymers by PARG. The suppression of the oxidative stress TRPM2-mediated Ca2+ response by PARP inhibitors suggests that the H2O2 effect involves the production of free ADPR (Fonfria et al. 2004). Additionally, PARP-deficient DT40 cells, which express TRPM2, exhibit no oxidative stress Ca2+ responses (Buelow et al. 2008). Interestingly, H2O2-mediated Ca2+ influx via TRPM2 is negatively affected when experiments are performed at room temperature and recovers at 37°C (Hermosura et al. 2008; Wilkinson et al. 2008). Moreover, it has been shown that temperatures >35°C may activate TRPM2 channels or facilitate their activation by ADPR or cADPR in rat insulinoma RIN-5F cells (Togashi et al. 2006). However, the mechanisms involved in the temperature-dependent regulation of TRPM2 remain to be explored in more detail.

TRPM2 currents are also regulated by cellular acidification (Du et al. 2009b; Starkus et al. 2010; Yang et al. 2010). TRPM2 currents are completely suppressed when exposing cells to external or internal pH of 5–6 (Du et al. 2009b; Starkus et al. 2010). Although both extracellular and intracellular mechanisms have been proposed to account for the proton-mediated regulation of TRPM2, the simplest interpretation that is compatible with the data reported so far would suggest that protons compete with Na+ and Ca2+ for channel permeation, and channel closure results from a competitive antagonism of protons at an intracellular Ca2+ binding site (Csanady, 2010; Starkus et al. 2010). The physiological relevance of this mechanism of TRPM2 regulation is unknown. However, it may be important both pathologically during tissue acidification caused by, for example, injury, inflammation, tumours, or ischaemia (Stekelenburg et al. 2008; Holzer, 2009; Stock & Schwab, 2009), as well as physiologically in highly acidic stores like lysosomes, where TRPM2 is also expressed and acts as a Ca2+ release channel.

The nucleotide adenosine monophosphate (AMP) can also antagonize TRPM2 gating by ADPR with an IC50 of 70 μm in HEK cells overexpressing TRPM2 channels (Kolisek et al. 2005) and 10 μm is sufficient to suppress endogenous channels in neutrophils (Lange et al. 2008). The antagonism is thought to result from a competition for the Nudix domain, which itself generates AMP from ADPR through enzymatic pyrophosphatase activity. This could potentially represent a physiological autoregulatory negative feedback mechanism of TRPM2 activity. It is additionally possible that AMP regulation of TRPM2 is less direct (Toth & Csanady, 2010) by involving, for example, AMP-dependent kinase modulation of TRPM2.

TRPM2 currents can also be inhibited pharmacologically by 8Br-cADPR (Kolisek et al. 2005; Beck et al. 2006), 8Br-ADPR (Partida-Sanchez et al. 2007), flufenamic acid (50–1000 μm) (Hill et al. 2004a), imidazole antifungal agents (clotrimazole and econazole; both at 3–30 μm) (Hill et al. 2004b), N-(p-amylcinnamoyl) (Kraft et al. 2006), anthranilic acid (ACA; 1.7 μm) and 2-aminoethoxydephenyl borate (2-APB; 1.2 μm) (Kraft et al. 2006; Togashi et al. 2008). However, most of these molecules are not very potent, nor do they exhibit high TRPM2 specificity.

While cytosolic ADPR appears to be the primary activator of TRPM2, the nucleotide also has extracellular effects. External application of ADPR increases [Ca2+]i through Ca2+ release from stores in TRPM2 expressing rat pancreatic β-cell line RIN-5F, an effect that is blocked by inhibiting phospholipase C (PLC) and IP3 (Ishii et al. 2006a). This observation led our group to investigate ADPR-induced Ca2+ pathways in pancreatic β-cells (Lange et al. 2009). Application of 100 μm ADPR to intact wild-type HEK293 cells, which do not express native TRPM2 channels, produced a transient Ca2+ signal independently of extracellular Ca2+ or the expression of TRPM2 in the plasma membrane of these cells. This response was blocked by suramin, suggesting that ADPR activates P2Y receptors. Both pathways, P2Y and TRPM2 dependent, were confirmed in the pancreatic β-cell line INS-1, where ADPR additionally activates adenosine receptors. Thus, ADPR appears to represent a multifunctional messenger that can activate membrane receptors extracellularly and also serve as an intracellular gating mechanism for TRPM2 residing in the plasma membrane and intracellular organelles.

Alternative splicing isoforms

The first TRPM2 variants were identified in the monocytic cell line HL-60 and in neutrophil granulocytes as two short proteins, the TRPM2 ΔN which is missing 20 residues (K538–Q557) in the N-terminus and the TRPM2 ΔC lacking 34 residues (T1292–L1325) in the C-terminus (Wehage et al. 2002). TRPM2 ΔC responds to H2O2 but not to ADPR, supporting the hypothesis of direct activation of TRPM2 by H2O2 (Wehage et al. 2002). In contrast, TRPM2 ΔN fails to respond to H2O2 (Wehage et al. 2002). However, a later study failed to confirm the C terminus as a locus for direct H2O2 activation (Kuhn & Luckhoff, 2004).

A third variant, TRPM2-S (short), contains only the N-terminus and the first two transmembrane segments and is generated by an additional stop codon between exons 16 and 17. This variant has been found in bone marrow, brain and pulmonary arteries and aorta (Zhang et al. 2003; Vazquez & Valverde, 2006; Yang et al. 2006; Hecquet & Malik, 2009) and may act as a dominant negative inhibitor of TRPM2 activity (Zhang et al. 2003).

A further isoform has been detected in human striatum, but not in mouse (Uemura et al. 2005). This variant, named striatum short protein (SSF)-TRPM2, contains the transmembrane domains and the C-terminal region, including the Nudix domain, but lacks the N-terminal 214 amino acid residues and retains H2O2-induced Ca2+ influx activity (Uemura et al. 2005). Consistent with this, SSF-TRPM2 localizes in the plasma membrane when expressed in HEK293 cells.

Computational analysis of altered methylation patterns of DNA regions with tumoral interest, has also detected two interesting TRPM2 transcripts in melanoma, the melanoma-enriched antisense TRPM2 transcript, TRPM2-AS, and a tumour-enriched TRPM2 transcript, TRPM2-TE (Orfanelli et al. 2008). Both transcripts seem to be up-regulated in malignant melanoma with little or no detectable levels in normal melanocytes (Orfanelli et al. 2008). Their transcription site appears to be in the intron 24 of the TRPM2 gene. In addition, there appears to exist a TRPM2-TE variant lacking part of exon 26 and the entire exon 27 (like the TRPM2 ΔC splice variant), resulting in the deletion of 34 amino acids from the Nudix domain but still containing the elements for the enzymatic activity (Orfanelli et al. 2008). The mechanisms of expression and functions of these transcripts remain to be characterized in more detail (Vazquez & Valverde, 2006). Nevertheless, the plethora of TRPM2 transcripts indicates that they might represent adaptive mechanisms for the cellular regulation of TRPM2 channels in different tissues, cell types and (patho)physiological circumstances.

TRPM2 and pathologies

TRPM2 has been associated with diseases such as bipolar disorder and type II diabetes. However, little is known about the TRPM2-dependent pathways that are affected in these pathologies. TRPM2 can mobilize Ca2+ from both extracellular and intracellular compartments and as such its biological significance is related to cellular functions that are regulated by [Ca2+]i. Since TRPM2 is most abundantly expressed in the brain, it is not surprising that TRPM2 has also been associated with CNS pathologies. Patients with bipolar disorders type I present high basal [Ca2+]i, and the chromosome region 21q22.3 harbours genes that confer susceptibility to this pathology, including TRPM2 (Xu et al. 2006, 2009). Although TRPM2 variants with a single amino substitution (e.g. Asp543Glu) have been detected in patients with bipolar disorder, the relevance of these variants in the pathogenesis of the disease remains to be elucidated. TRPM2 has also been linked to amyotrophic lateral sclerosis and parkinsonism–dementia (Hermosura & Garruto, 2007). Here, a TRPM2 mutation (P1018L) results in channels that inactivate more rapidly than wild-type channels, resulting in reduced Ca2+ entry. Again, the cellular and functional context of TRPM2 in these pathologies remains to be demonstrated.

In addition to the CNS, TRPM2 has also been identified in pancreatic β-cells Qian et al. 2002; Ishii et al. 2006a,b; Togashi et al. 2006; Bari et al. 2009; Lange et al. 2009). Moreover, TRPM2 activity has been linked to insulin secretion and alloxan- and H2O2-mediated apoptosis of insulin-secreting cells, suggesting a role of TRPM2 in diabetes (Herson & Ashford, 1997; Herson et al. 1999; Togashi et al. 2006; Uchida et al. 2010). A recent study using the TRPM2 KO mouse model revealed involvement of this channel in insulin secretion in mouse pancreatic β-cells (Uchida et al. 2010). In this study, basal blood glucose levels were higher in TRPM2-KO mice than in WT mice, while plasma insulin levels were similar. In isolated β-cells, TRPM2-KO cells produced smaller Ca2+ signals in response to high concentrations of glucose and incretin hormone than WT cells, resulting in reduced insulin secretion from pancreatic islets of TRPM2-KO mice. These results indicate that TRPM2 is involved in insulin secretion stimulated by glucose. In contrast, a clinical study of a patient population with type 2 diabetes mellitus found no association of the genetic TRPM2 variants rs2838553, rs2838554, rs4818917, rs1619968, rs1785452, rs2238725, rs2010779, rs9979491 and rs1573477 with this chronic disease (Romero et al. 2010). However, the variants rs2838553, rs2838554 and rs4818917 showed negative association with a homeostatic model assessment of β-cell function, which determines insulin resistance and β-cell function, hinting at the possibility that TRPM2 activity might regulate β-cell function. Further studies examining other variants are necessary to establish a role of TRPM2 in diabetes.

TRPM2 has also been identified in numerous cell types of the immune system, including neutrophils, monocytes, macrophages and lymphocytes. The ability of H2O2 to activate TRPM2 channels has attracted interest to this channel as a potential mechanism for pathogenic processes that are characterized by an increased oxidative microenvironment, including carcinogenesis, inflammation, ischaemia–reperfusion injury, neurodegenerative disorders, diabetes and others. Endogenously, H2O2 is generated through oxidative phosphorylation in mitochondria (Roede & Jones, 2010). Exogenously, H2O2 generation is induced in concert with other reactive oxygen species (ROS), such as superperoxides (O2) and hydroxyl radicals (OH.), as a result of exposure to a wide variety of external factors including certain drugs, pollutants, heavy metals, heat, UV or visible light, and other forms of ionizing radiation (Cheeseman & Slater, 1993). Uncontrolled generation of ROS causes significant damage to a wide range of biological molecules such as DNA, lipids, proteins, carbohydrates or any nearby molecule causing a cascade of chain reactions resulting in cellular damage and disease (Akyol et al. 2002). H2O2 is an oxidizing agent, and although it is not especially reactive, it is a source of OH in the presence of reactive transition metal ions, which is an extremely reactive oxidizing radical that affects most biomolecules (Naziroglu, 2007).

H2O2 also functions as an antimicrobial agent in immune cells and as a signal-transduction molecule (Droge, 2002). Importantly, H2O2 activates TRPM2 and in cells that express TRPM2 at high levels, this can induce cell death by sustained increases in [Ca2+]i (Hara et al. 2002). The mechanism by which H2O2 activates TRPM2 channels is not entirely resolved. The slow kinetics of TRPM2 activation by external application of H2O2 suggests that H2O2 needs to cross the plasma membrane and generate ADPR or other factors to modulate TRPM2. Mutations in the Nudix domain or treatment with PARP inhibitors prevents or strongly impairs TRPM2 activation by H2O2 (Fonfria et al. 2004). [Ca2+]i increases following external H2O2 application are also lost in PARP-deficient DT40 cells (Buelow et al. 2008).

The significance of the TRPM2 channel in processes that depend on oxidative stress was recently confirmed for dextran sulfate sodium-induced colitis in mice, a model for human inflammatory bowel disease (Yamamoto et al. 2008). TRPM2 KO mice were protected from tissue damage due to reduced production of CXCL2 chemokines by monocytes, resulting in diminished recruitment of neutrophils to the site of inflammation. Although neutrophils also express TRPM2, the chemotactic response to CXCL2 was not perturbed in TRPM2-deficient mice. Therefore, impaired neutrophil infiltration to the site of injury in TRPM2KO mice was a consequence of a defective production of CXCL2 in monocytes or macrophages (Yamamoto et al. 2008). The impaired chemokine production in cells lacking TRPM2 was linked to a defect in TRPM2-mediated Ca2+ influx, which normally proceeds via activation of the Ca2+-dependent kinase Pyk2 and activation of the Erk/NFkB pathway (Yamamoto et al. 2008, 2010). The suppression of H2O2-mediated Ca2+ signalling and CXCL8 (the human equivalent of CXCL2) production through the Erk pathway was also confirmed in the human monocytic cell line U-937 (Yamamoto et al. 2008).

Conclusions

Since the discovery of TRPM2 a decade ago, there have been many important findings that have advanced the understanding of its protein structure and the mechanisms that regulate the activity of this channel. There have also been important advances in the understanding at the gene expression level and most recently its cellular localization and function. This channel has revealed many fascinating features that make it one of the most versatile and intriguing ion channels. Future work will focus on mechanisms that regulate TRPM2 abundance in the various cellular compartments it can be expressed in and what factors determine intracellular versus plasma membrane localization. Given the highly integrative nature of this channel, it is likely that additional cellular regulatory factors will be discovered. Further investigations will focus on signalling pathways upstream of TRPM2, so as to determine the cellular mechanisms that mobilize ADPR under physiological and pathological conditions. Finally, it will be interesting to learn to what extent and in what role TRPM2 participates in pathophysiological conditions, both in the CNS and in peripheral tissues. Hopefully, the discovery of pharmacological tools will aid in characterizing TRPM2, with the prospect of using some of these tools therapeutically.

References

  1. Akyol O, Herken H, Uz E, Fadillioglu E, Unal S, Sogut S, Ozyurt H, Savas HA. The indices of endogenous oxidative and antioxidative processes in plasma from schizophrenic patients. The possible role of oxidant/antioxidant imbalance. Prog Neuropsychopharmacol Biol Psychiatry. 2002;26:995–1005. doi: 10.1016/s0278-5846(02)00220-8. [DOI] [PubMed] [Google Scholar]
  2. Ayub K, Hallett MB. The mitochondrial ADPR link between Ca2+ store release and Ca2+ influx channel opening in immune cells. FASEB J. 2004;18:1335–1338. doi: 10.1096/fj.04-1888hyp. [DOI] [PubMed] [Google Scholar]
  3. Bari MR, Akbar S, Eweida M, Kuhn FJ, Gustafsson AJ, Luckhoff A, Islam MS. H2O2-induced Ca2+ influx and its inhibition by N-(p-amylcinnamoyl) anthranilic acid in the β-cells: involvement of TRPM2 channels. J Cell Mol Med. 2009;13:3260–3267. doi: 10.1111/j.1582-4934.2009.00737.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Beck A, Kolisek M, Bagley LA, Fleig A, Penner R. Nicotinic acid adenine dinucleotide phosphate and cyclic ADP-ribose regulate TRPM2 channels in T lymphocytes. FASEB J. 2006;20:962–964. doi: 10.1096/fj.05-5538fje. [DOI] [PubMed] [Google Scholar]
  5. Bessman MJ, Frick DN, O'Handley SF. The MutT proteins or “Nudix” hydrolases, a family of versatile, widely distributed, “housecleaning” enzymes. J Biol Chem. 1996;271:25059–25062. doi: 10.1074/jbc.271.41.25059. [DOI] [PubMed] [Google Scholar]
  6. Bezzerides VJ, Ramsey IS, Kotecha S, Greka A, Clapham DE. Rapid vesicular translocation and insertion of TRP channels. Nat Cell Biol. 2004;6:709–720. doi: 10.1038/ncb1150. [DOI] [PubMed] [Google Scholar]
  7. Buelow B, Song Y, Scharenberg AM. The Poly(ADP-ribose) polymerase PARP-1 is required for oxidative stress-induced TRPM2 activation in lymphocytes. J Biol Chem. 2008;283:24571–24583. doi: 10.1074/jbc.M802673200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Caiafa P, Guastafierro T, Zampieri M. Epigenetics: poly(ADP-ribosyl)ation of PARP-1 regulates genomic methylation patterns. FASEB J. 2009;23:672–678. doi: 10.1096/fj.08-123265. [DOI] [PubMed] [Google Scholar]
  9. Carter RN, Tolhurst G, Walmsley G, Vizuete-Forster M, Miller N, Mahaut-Smith MP. Molecular and electrophysiological characterization of transient receptor potential ion channels in the primary murine megakaryocyte. J Physiol. 2006;576:151–162. doi: 10.1113/jphysiol.2006.113886. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Cheeseman KH, Slater TF. An introduction to free radical biochemistry. Br Med Bull. 1993;49:481–493. doi: 10.1093/oxfordjournals.bmb.a072625. [DOI] [PubMed] [Google Scholar]
  11. Cheng X, Shen D, Samie M, Xu H. Mucolipins: Intracellular TRPML1–3 channels. FEBS Lett. 2010;584:2013–2021. doi: 10.1016/j.febslet.2009.12.056. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Cosens DJ, Manning A. Abnormal electroretinogram from a Drosophila mutant. Nature. 1969;224:285–287. doi: 10.1038/224285a0. [DOI] [PubMed] [Google Scholar]
  13. Csanady L. Permeating proton found guilty in compromising TRPM2 channel activity. J Physiol. 2010;588:1661–1662. doi: 10.1113/jphysiol.2010.190223. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Csanady L, Torocsik B. Four Ca2+ ions activate TRPM2 channels by binding in deep crevices near the pore but intracellularly of the gate. J Gen Physiol. 2009;133:189–203. doi: 10.1085/jgp.200810109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Dong XP, Wang X, Xu H. TRP channels of intracellular membranes. J Neurochem. 2010;113:313–328. doi: 10.1111/j.1471-4159.2010.06626.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Droge W. Free radicals in the physiological control of cell function. Physiol Rev. 2002;82:47–95. doi: 10.1152/physrev.00018.2001. [DOI] [PubMed] [Google Scholar]
  17. Du J, Xie J, Yue L. Intracellular calcium activates TRPM2 and its alternative spliced isoforms. Proc Natl Acad Sci U S A. 2009a;106:7239–7244. doi: 10.1073/pnas.0811725106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Du J, Xie J, Yue L. Modulation of TRPM2 by acidic pH and the underlying mechanisms for pH sensitivity. J Gen Physiol. 2009b;134:471–488. doi: 10.1085/jgp.200910254. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Esposito E, Cuzzocrea S. Superoxide, NO, peroxynitrite and PARP in circulatory shock and inflammation. Front Biosci. 2009;14:263–296. doi: 10.2741/3244. [DOI] [PubMed] [Google Scholar]
  20. Fauzee NJ, Pan J, Wang YL. PARP and PARG inhibitors: new therapeutic targets in cancer treatment. Pathol Oncol Res. 2010;16:469–478. doi: 10.1007/s12253-010-9266-6. [DOI] [PubMed] [Google Scholar]
  21. Fleig A, Penner R. Emerging roles of TRPM channels. Novartis Found Symp. 2004a;258:248–258. discussion 258–266. [PubMed] [Google Scholar]
  22. Fleig A, Penner R. The TRPM ion channel subfamily: molecular, biophysical and functional features. Trends Pharmacol Sci. 2004b;25:633–639. doi: 10.1016/j.tips.2004.10.004. [DOI] [PubMed] [Google Scholar]
  23. Fonfria E, Marshall IC, Benham CD, Boyfield I, Brown JD, Hill K, Hughes JP, Skaper SD, McNulty S. TRPM2 channel opening in response to oxidative stress is dependent on activation of poly(ADP-ribose) polymerase. Br J Pharmacol. 2004;143:186–192. doi: 10.1038/sj.bjp.0705914. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Gasser A, Glassmeier G, Fliegert R, Langhorst MF, Meinke S, Hein D, Kruger S, Weber K, Heiner I, Oppenheimer N, Schwarz JR, Guse AH. Activation of T cell calcium influx by the second messenger ADP-ribose. J Biol Chem. 2006;281:2489–2496. doi: 10.1074/jbc.M506525200. [DOI] [PubMed] [Google Scholar]
  25. Gaudet R. TRP channels entering the structural era. J Physiol. 2008;586:3565–3575. doi: 10.1113/jphysiol.2008.155812. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Gaudet R. Divide and conquer: high resolution structural information on TRP channel fragments. J Gen Physiol. 2009;133:231–237. doi: 10.1085/jgp.200810137. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Gees M, Colsoul B, Nilius B. The role of transient receptor potential cation channels in Ca2+ signaling. Cold Spring Harb Perspect Biol. 2010;2:a003962. doi: 10.1101/cshperspect.a003962. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Goel M, Sinkins WG, Schilling WP. Selective association of TRPC channel subunits in rat brain synaptosomes. J Biol Chem. 2002;277:48303–48310. doi: 10.1074/jbc.M207882200. [DOI] [PubMed] [Google Scholar]
  29. Grubisha O, Rafty LA, Takanishi CL, Xu X, Tong L, Perraud AL, Scharenberg AM, Denu JM. Metabolite of SIR2 reaction modulates TRPM2 ion channel. J Biol Chem. 2006;281:14057–14065. doi: 10.1074/jbc.M513741200. [DOI] [PubMed] [Google Scholar]
  30. Hara Y, Wakamori M, Ishii M, Maeno E, Nishida M, Yoshida T, Yamada H, Shimizu S, Mori E, Kudoh J, Shimizu N, Kurose H, Okada Y, Imoto K, Mori Y. LTRPC2 Ca2+-permeable channel activated by changes in redox status confers susceptibility to cell death. Mol Cell. 2002;9:163–173. doi: 10.1016/s1097-2765(01)00438-5. [DOI] [PubMed] [Google Scholar]
  31. Hecquet CM, Ahmmed GU, Malik AB. TRPM2 channel regulates endothelial barrier function. Adv Exp Med Biol. 2010;661:155–167. doi: 10.1007/978-1-60761-500-2_10. [DOI] [PubMed] [Google Scholar]
  32. Hecquet CM, Ahmmed GU, Vogel SM, Malik AB. Role of TRPM2 channel in mediating H2O2-induced Ca2+ entry and endothelial hyperpermeability. Circ Res. 2008;102:347–355. doi: 10.1161/CIRCRESAHA.107.160176. [DOI] [PubMed] [Google Scholar]
  33. Hecquet CM, Malik AB. Role of H2O2-activated TRPM2 calcium channel in oxidant-induced endothelial injury. Thromb Haemost. 2009;101:619–625. [PMC free article] [PubMed] [Google Scholar]
  34. Heiner I, Eisfeld J, Luckhoff A. Role and regulation of TRP channels in neutrophil granulocytes. Cell Calcium. 2003;33:533–540. doi: 10.1016/s0143-4160(03)00058-7. [DOI] [PubMed] [Google Scholar]
  35. Hermosura MC, Cui AM, Go RC, Davenport B, Shetler CM, Heizer JW, Schmitz C, Mocz G, Garruto RM, Perraud AL. Altered functional properties of a TRPM2 variant in Guamanian ALS and PD. Proc Natl Acad Sci U S A. 2008;105:18029–18034. doi: 10.1073/pnas.0808218105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Hermosura MC, Garruto RM. TRPM7 and TRPM2-candidate susceptibility genes for Western Pacific ALS and PD? Biochim Biophys Acta. 2007;1772:822–835. doi: 10.1016/j.bbadis.2007.02.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Herson PS, Ashford ML. Activation of a novel non-selective cation channel by alloxan and H2O2 in the rat insulin-secreting cell line CRI-G1. J Physiol. 1997;501:59–66. doi: 10.1111/j.1469-7793.1997.059bo.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Herson PS, Lee K, Pinnock RD, Hughes J, Ashford ML. Hydrogen peroxide induces intracellular calcium overload by activation of a non-selective cation channel in an insulin-secreting cell line. J Biol Chem. 1999;274:833–841. doi: 10.1074/jbc.274.2.833. [DOI] [PubMed] [Google Scholar]
  39. Hill K, Benham CD, McNulty S, Randall AD. Flufenamic acid is a pH-dependent antagonist of TRPM2 channels. Neuropharmacology. 2004a;47:450–460. doi: 10.1016/j.neuropharm.2004.04.014. [DOI] [PubMed] [Google Scholar]
  40. Hill K, McNulty S, Randall AD. Inhibition of TRPM2 channels by the antifungal agents clotrimazole and econazole. Naunyn Schmiedebergs Arch Pharmacol. 2004b;370:227–237. doi: 10.1007/s00210-004-0981-y. [DOI] [PubMed] [Google Scholar]
  41. Hill K, Tigue NJ, Kelsell RE, Benham CD, McNulty S, Schaefer M, Randall AD. Characterisation of recombinant rat TRPM2 and a TRPM2-like conductance in cultured rat striatal neurones. Neuropharmacology. 2006;50:89–97. doi: 10.1016/j.neuropharm.2005.08.021. [DOI] [PubMed] [Google Scholar]
  42. Hofmann T, Schaefer M, Schultz G, Gudermann T. Subunit composition of mammalian transient receptor potential channels in living cells. Proc Natl Acad Sci U S A. 2002;99:7461–7466. doi: 10.1073/pnas.102596199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Holzer P. Acid-sensitive ion channels and receptors. Handb Exp Pharmacol. 2009:283–332. doi: 10.1007/978-3-540-79090-7_9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Inamura K, Sano Y, Mochizuki S, Yokoi H, Miyake A, Nozawa K, Kitada C, Matsushime H, Furuichi K. Response to ADP-ribose by activation of TRPM2 in the CRI-G1 insulinoma cell line. J Membr Biol. 2003;191:201–207. doi: 10.1007/s00232-002-1057-x. [DOI] [PubMed] [Google Scholar]
  45. Ishii M, Shimizu S, Hagiwara T, Wajima T, Miyazaki A, Mori Y, Kiuchi Y. Extracellular-added ADP-ribose increases intracellular free Ca2+ concentration through Ca2+ release from stores, but not through TRPM2-mediated Ca2+ entry, in rat β-cell line RIN-5F. J Pharmacol Sci. 2006a;101:174–178. doi: 10.1254/jphs.scj06001x. [DOI] [PubMed] [Google Scholar]
  46. Ishii M, Shimizu S, Hara Y, Hagiwara T, Miyazaki A, Mori Y, Kiuchi Y. Intracellular-produced hydroxyl radical mediates H2O2-induced Ca2+ influx and cell death in rat β-cell line RIN-5F. Cell Calcium. 2006b;39:487–494. doi: 10.1016/j.ceca.2006.01.013. [DOI] [PubMed] [Google Scholar]
  47. Jiang LH. Subunit interaction in channel assembly and functional regulation of transient receptor potential melastatin (TRPM) channels. Biochem Soc Trans. 2007;35:86–88. doi: 10.1042/BST0350086. [DOI] [PubMed] [Google Scholar]
  48. Jiang LH, Yang W, Zou J, Beech DJ. TRPM2 channel properties, functions and therapeutic potentials. Expert Opin Ther Targets. 2010;14:973–988. doi: 10.1517/14728222.2010.510135. [DOI] [PubMed] [Google Scholar]
  49. Kolisek M, Beck A, Fleig A, Penner R. Cyclic ADP-ribose and hydrogen peroxide synergize with ADP-ribose in the activation of TRPM2 channels. Mol Cell. 2005;18:61–69. doi: 10.1016/j.molcel.2005.02.033. [DOI] [PubMed] [Google Scholar]
  50. Kraft R, Grimm C, Frenzel H, Harteneck C. Inhibition of TRPM2 cation channels by N-(p-amylcinnamoyl) anthranilic acid. Br J Pharmacol. 2006;148:264–273. doi: 10.1038/sj.bjp.0706739. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Kraft R, Grimm C, Grosse K, Hoffmann A, Sauerbruch S, Kettenmann H, Schultz G, Harteneck C. Hydrogen peroxide and ADP-ribose induce TRPM2-mediated calcium influx and cation currents in microglia. Am J Physiol Cell Physiol. 2004;286:C129–137. doi: 10.1152/ajpcell.00331.2003. [DOI] [PubMed] [Google Scholar]
  52. Kuhn FJ, Luckhoff A. Sites of the NUDT9-H domain critical for ADP-ribose activation of the cation channel TRPM2. J Biol Chem. 2004;279:46431–46437. doi: 10.1074/jbc.M407263200. [DOI] [PubMed] [Google Scholar]
  53. Lange I, Penner R, Fleig A, Beck A. Synergistic regulation of endogenous TRPM2 channels by adenine dinucleotides in primary human neutrophils. Cell Calcium. 2008;44:604–615. doi: 10.1016/j.ceca.2008.05.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Lange I, Yamamoto S, Partida-Sanchez S, Mori Y, Fleig A, Penner R. TRPM2 functions as a lysosomal Ca2+-release channel in beta cells. Sci Signal. 2009;2:ra23. doi: 10.1126/scisignal.2000278. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Latorre R, Zaelzer C, Brauchi S. Structure-functional intimacies of transient receptor potential channels. Q Rev Biophys. 2009;42:201–246. doi: 10.1017/S0033583509990072. [DOI] [PubMed] [Google Scholar]
  56. Li M, Jiang J, Yue L. Functional characterization of homo- and heteromeric channel kinases TRPM6 and TRPM7. J Gen Physiol. 2006;127:525–537. doi: 10.1085/jgp.200609502. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Lund F, Solvason N, Grimaldi JC, Parkhouse RM, Howard M. Murine CD38: an immunoregulatory ectoenzyme. Immunol Today. 1995;16:469–473. doi: 10.1016/0167-5699(95)80029-8. [DOI] [PubMed] [Google Scholar]
  58. Lund FE. Signaling properties of CD38 in the mouse immune system: enzyme-dependent and -independent roles in immunity. Mol Med. 2006;12:328–333. doi: 10.2119/2006-00099.Lund. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Lund FE, Cockayne DA, Randall TD, Solvason N, Schuber F, Howard MC. CD38: a new paradigm in lymphocyte activation and signal transduction. Immunol Rev. 1998;161:79–93. doi: 10.1111/j.1600-065x.1998.tb01573.x. [DOI] [PubMed] [Google Scholar]
  60. Malavasi F, Deaglio S, Ferrero E, Funaro A, Sancho J, Ausiello CM, Ortolan E, Vaisitti T, Zubiaur M, Fedele G, Aydin S, Tibaldi EV, Durelli I, Lusso R, Cozno F, Horenstein AL. CD38 and CD157 as receptors of the immune system: a bridge between innate and adaptive immunity. Mol Med. 2006;12:334–341. doi: 10.2119/2006-00094.Malavasi. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. McHugh D, Flemming R, Xu SZ, Perraud AL, Beech DJ. Critical intracellular Ca2+ dependence of transient receptor potential melastatin 2 (TRPM2) cation channel activation. J Biol Chem. 2003;278:11002–11006. doi: 10.1074/jbc.M210810200. [DOI] [PubMed] [Google Scholar]
  62. Montell C, Rubin GM. Molecular characterization of the Drosophila trp locus: a putative integral membrane protein required for phototransduction. Neuron. 1989;2:1313–1323. doi: 10.1016/0896-6273(89)90069-x. [DOI] [PubMed] [Google Scholar]
  63. Morenilla-Palao C, Planells-Cases R, Garcia-Sanz N, Ferrer-Montiel A. Regulated exocytosis contributes to protein kinase C potentiation of vanilloid receptor activity. J Biol Chem. 2004;279:25665–25672. doi: 10.1074/jbc.M311515200. [DOI] [PubMed] [Google Scholar]
  64. Nagamine K, Kudoh J, Minoshima S, Kawasaki K, Asakawa S, Ito F, Shimizu N. Molecular cloning of a novel putative Ca2+ channel protein (TRPC7) highly expressed in brain. Genomics. 1998;54:124–131. doi: 10.1006/geno.1998.5551. [DOI] [PubMed] [Google Scholar]
  65. Naziroglu M. New molecular mechanisms on the activation of TRPM2 channels by oxidative stress and ADP-ribose. Neurochem Res. 2007;32:1990–2001. doi: 10.1007/s11064-007-9386-x. [DOI] [PubMed] [Google Scholar]
  66. Naziroglu M, Luckhoff A. A calcium influx pathway regulated separately by oxidative stress and ADP-ribose in TRPM2 channels: single channel events. Neurochem Res. 2008;33:1256–1262. doi: 10.1007/s11064-007-9577-5. [DOI] [PubMed] [Google Scholar]
  67. Nilius B. Transient receptor potential (TRP) cation channels: rewarding unique proteins. Bull Mem Acad R Med Belg. 2007;162:244–253. [PubMed] [Google Scholar]
  68. Oancea E, Wolfe JT, Clapham DE. Functional TRPM7 channels accumulate at the plasma membrane in response to fluid flow. Circ Res. 2006;98:245–253. doi: 10.1161/01.RES.0000200179.29375.cc. [DOI] [PubMed] [Google Scholar]
  69. Olah ME, Jackson MF, Li H, Perez Y, Sun HS, Kiyonaka S, Mori Y, Tymianski M, MacDonald JF. Ca2+-dependent induction of TRPM2 currents in hippocampal neurons. J Physiol. 2009;587:965–979. doi: 10.1113/jphysiol.2008.162289. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Orfanelli U, Wenke AK, Doglioni C, Russo V, Bosserhoff AK, Lavorgna G. Identification of novel sense and antisense transcription at the TRPM2 locus in cancer. Cell Res. 2008;18:1128–1140. doi: 10.1038/cr.2008.296. [DOI] [PubMed] [Google Scholar]
  71. Partida-Sanchez S, Gasser A, Fliegert R, Siebrands CC, Dammermann W, Shi G, Mousseau BJ, Sumoza-Toledo A, Bhagat H, Walseth TF, Guse AH, Lund FE. Chemotaxis of mouse bone marrow neutrophils and dendritic cells is controlled by ADP-ribose, the major product generated by the CD38 enzyme reaction. J Immunol. 2007;179:7827–7839. doi: 10.4049/jimmunol.179.11.7827. [DOI] [PubMed] [Google Scholar]
  72. Partida-Sanchez S, Iribarren P, Moreno-Garcia ME, Gao JL, Murphy PM, Oppenheimer N, Wang JM, Lund FE. Chemotaxis and calcium responses of phagocytes to formyl peptide receptor ligands is differentially regulated by cyclic ADP ribose. J Immunol. 2004;172:1896–1906. doi: 10.4049/jimmunol.172.3.1896. [DOI] [PubMed] [Google Scholar]
  73. Partida-Sanchez S, Randall TD, Lund FE. Innate immunity is regulated by CD38, an ecto-enzyme with ADP-ribosyl cyclase activity. Microbes Infect. 2003;5:49–58. doi: 10.1016/s1286-4579(02)00055-2. [DOI] [PubMed] [Google Scholar]
  74. Perraud AL, Fleig A, Dunn CA, Bagley LA, Launay P, Schmitz C, Stokes AJ, Zhu Q, Bessman MJ, Penner R, Kinet JP, Scharenberg AM. ADP-ribose gating of the calcium-permeable LTRPC2 channel revealed by Nudix motif homology. Nature. 2001;411:595–599. doi: 10.1038/35079100. [DOI] [PubMed] [Google Scholar]
  75. Perraud AL, Shen B, Dunn CA, Rippe K, Smith MK, Bessman MJ, Stoddard BL, Scharenberg AM. NUDT9, a member of the Nudix hydrolase family, is an evolutionarily conserved mitochondrial ADP-ribose pyrophosphatase. J Biol Chem. 2003a;278:1794–1801. doi: 10.1074/jbc.M205601200. [DOI] [PubMed] [Google Scholar]
  76. Perraud AL, Shen B, Dunn CA, Rippe K, Smith MK, Bessman MJ, Stoddard BL, Scharenberg AM. NUDT9, a member of the Nudix hydrolase family, is an evolutionarily conserved mitochondrial ADP-ribose pyrophosphatase. J Biol Chem. 2003b;278:1794–1801. doi: 10.1074/jbc.M205601200. [DOI] [PubMed] [Google Scholar]
  77. Perraud AL, Takanishi CL, Shen B, Kang S, Smith MK, Schmitz C, Knowles HM, Ferraris D, Li W, Zhang J, Stoddard BL, Scharenberg AM. Accumulation of free ADP-ribose from mitochondria mediates oxidative stress-induced gating of TRPM2 cation channels. J Biol Chem. 2005;280:6138–6148. doi: 10.1074/jbc.M411446200. [DOI] [PubMed] [Google Scholar]
  78. Piper RC, Luzio JP. CUPpling calcium to lysosomal biogenesis. Trends Cell Biol. 2004;14:471–473. doi: 10.1016/j.tcb.2004.07.010. [DOI] [PubMed] [Google Scholar]
  79. Qian F, Huang P, Ma L, Kuznetsov A, Tamarina N, Philipson LH. TRP genes: candidates for nonselective cation channels and store-operated channels in insulin-secreting cells. Diabetes. 2002;51(Suppl 1):S183–189. doi: 10.2337/diabetes.51.2007.s183. [DOI] [PubMed] [Google Scholar]
  80. Ramsey IS, Delling M, Clapham DE. An introduction to TRP channels. Annu Rev Physiol. 2006;68:619–647. doi: 10.1146/annurev.physiol.68.040204.100431. [DOI] [PubMed] [Google Scholar]
  81. Roede JR, Jones DP. Reactive species and mitochondrial dysfunction: mechanistic significance of 4-hydroxynonenal. Environ Mol Mutagen. 2010;51:380–390. doi: 10.1002/em.20553. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Romero JR, Germer S, Castonguay AJ, Barton NS, Martin M, Zee RY. Gene variation of the transient receptor potential cation channel, subfamily M, member 2 (TRPM2) and type 2 diabetes mellitus: A case-control study. Clin Chim Acta. 2010;411:1437–1440. doi: 10.1016/j.cca.2010.05.036. [DOI] [PubMed] [Google Scholar]
  83. Sano Y, Inamura K, Miyake A, Mochizuki S, Yokoi H, Matsushime H, Furuichi K. Immunocyte Ca2+ influx system mediated by LTRPC2. Science. 2001;293:1327–1330. doi: 10.1126/science.1062473. [DOI] [PubMed] [Google Scholar]
  84. Singh BB, Lockwich TP, Bandyopadhyay BC, Liu X, Bollimuntha S, Brazer SC, Combs C, Das S, Leenders AG, Sheng ZH, Knepper MA, Ambudkar SV, Ambudkar IS. VAMP2-dependent exocytosis regulates plasma membrane insertion of TRPC3 channels and contributes to agonist-stimulated Ca2+ influx. Mol Cell. 2004;15:635–646. doi: 10.1016/j.molcel.2004.07.010. [DOI] [PubMed] [Google Scholar]
  85. Song MY, Yuan JX. Introduction to TRP channels: structure, function, and regulation. Adv Exp Med Biol. 2010;661:99–108. doi: 10.1007/978-1-60761-500-2_6. [DOI] [PubMed] [Google Scholar]
  86. Starkus J, Beck A, Fleig A, Penner R. Regulation of TRPM2 by extra- and intracellular calcium. J Gen Physiol. 2007;130:427–440. doi: 10.1085/jgp.200709836. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Starkus JG, Fleig A, Penner R. The calcium-permeable non-selective cation channel TRPM2 is modulated by cellular acidification. J Physiol. 2010;588:1227–1240. doi: 10.1113/jphysiol.2010.187476. [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Stekelenburg A, Gawlitta D, Bader DL, Oomens CW. Deep tissue injury: how deep is our understanding? Arch Phys Med Rehabil. 2008;89:1410–1413. doi: 10.1016/j.apmr.2008.01.012. [DOI] [PubMed] [Google Scholar]
  89. Stock C, Schwab A. Protons make tumor cells move like clockwork. Pflugers Arch. 2009;458:981–992. doi: 10.1007/s00424-009-0677-8. [DOI] [PubMed] [Google Scholar]
  90. Talavera K, Nilius B, Voets T. Neuronal TRP channels: thermometers, pathfinders and life-savers. Trends Neurosci. 2008;31:287–295. doi: 10.1016/j.tins.2008.03.002. [DOI] [PubMed] [Google Scholar]
  91. Thebault S, Lemonnier L, Bidaux G, Flourakis M, Bavencoffe A, Gordienko D, Roudbaraki M, Delcourt P, Panchin Y, Shuba Y, Skryma R, Prevarskaya N. Novel role of cold/menthol-sensitive transient receptor potential melastatine family member 8 (TRPM8) in the activation of store-operated channels in LNCaP human prostate cancer epithelial cells. J Biol Chem. 2005;280:39423–39435. doi: 10.1074/jbc.M503544200. [DOI] [PubMed] [Google Scholar]
  92. Togashi K, Hara Y, Tominaga T, Higashi T, Konishi Y, Mori Y, Tominaga M. TRPM2 activation by cyclic ADP-ribose at body temperature is involved in insulin secretion. EMBO J. 2006;25:1804–1815. doi: 10.1038/sj.emboj.7601083. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Togashi K, Inada H, Tominaga M. Inhibition of the transient receptor potential cation channel TRPM2 by 2-aminoethoxydiphenyl borate (2-APB) Br J Pharmacol. 2008;153:1324–1330. doi: 10.1038/sj.bjp.0707675. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Tong L, Denu JM. Function and metabolism of sirtuin metabolite O-acetyl-ADP-ribose. Biochim Biophys Acta. 2010;1804:1617–1625. doi: 10.1016/j.bbapap.2010.02.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Tong Q, Zhang W, Conrad K, Mostoller K, Cheung JY, Peterson BZ, Miller BA. Regulation of the transient receptor potential channel TRPM2 by the Ca2+ sensor calmodulin. J Biol Chem. 2006;281:9076–9085. doi: 10.1074/jbc.M510422200. [DOI] [PubMed] [Google Scholar]
  96. Toth B, Csanady L. Identification of direct and indirect effectors of the transient receptor potential melastatin 2 (TRPM2) cation channel. J Biol Chem. 2010;285:30091–30102. doi: 10.1074/jbc.M109.066464. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Uchida K, Dezaki K, Damdindorj B, Inada H, Shiuchi T, Mori Y, Yada T, Minokoshi Y, Tominaga M. Lack of TRPM2 impaired insulin secretion and glucose metabolisms in mice. Diabetes. 2010 doi: 10.2337/db10-0276. in press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Uemura T, Kudoh J, Noda S, Kanba S, Shimizu N. Characterization of human and mouse TRPM2 genes: identification of a novel N-terminal truncated protein specifically expressed in human striatum. Biochem Biophys Res Commun. 2005;328:1232–1243. doi: 10.1016/j.bbrc.2005.01.086. [DOI] [PubMed] [Google Scholar]
  99. Vazquez E, Valverde MA. A review of TRP channels splicing. Semin Cell Dev Biol. 2006;17:607–617. doi: 10.1016/j.semcdb.2006.11.004. [DOI] [PubMed] [Google Scholar]
  100. Venkatachalam K, Montell C. TRP channels. Annu Rev Biochem. 2007;76:387–417. doi: 10.1146/annurev.biochem.75.103004.142819. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Wehage E, Eisfeld J, Heiner I, Jungling E, Zitt C, Luckhoff A. Activation of the cation channel long transient receptor potential channel 2 (LTRPC2) by hydrogen peroxide. A splice variant reveals a mode of activation independent of ADP-ribose. J Biol Chem. 2002;277:23150–23156. doi: 10.1074/jbc.M112096200. [DOI] [PubMed] [Google Scholar]
  102. Wilkinson JA, Scragg JL, Boyle JP, Nilius B, Peers C. H2O2-stimulated Ca2+ influx via TRPM2 is not the sole determinant of subsequent cell death. Pflugers Arch. 2008;455:1141–1151. doi: 10.1007/s00424-007-0384-2. [DOI] [PubMed] [Google Scholar]
  103. Xu C, Li PP, Cooke RG, Parikh SV, Wang K, Kennedy JL, Warsh JJ. TRPM2 variants and bipolar disorder risk: confirmation in a family-based association study. Bipolar Disord. 2009;11:1–10. doi: 10.1111/j.1399-5618.2008.00655.x. [DOI] [PubMed] [Google Scholar]
  104. Xu C, Macciardi F, Li PP, Yoon IS, Cooke RG, Hughes B, Parikh SV, McIntyre RS, Kennedy JL, Warsh JJ. Association of the putative susceptibility gene, transient receptor potential protein melastatin type 2, with bipolar disorder. Am J Med Genet B Neuropsychiatr Genet. 2006;141B:36–43. doi: 10.1002/ajmg.b.30239. [DOI] [PubMed] [Google Scholar]
  105. Yamamoto S, Shimizu S, Kiyonaka S, Takahashi N, Wajima T, Hara Y, Negoro T, Hiroi T, Kiuchi Y, Okada T, Kaneko S, Lange I, Fleig A, Penner R, Nishi M, Takeshima H, Mori Y. TRPM2-mediated Ca2+ influx induces chemokine production in monocytes that aggravates inflammatory neutrophil infiltration. Nat Med. 2008;14:738–747. doi: 10.1038/nm1758. [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Yamamoto S, Takahashi N, Mori Y. Chemical physiology of oxidative stress-activated TRPM2 and TRPC5 channels. Prog Biophys Mol Biol. 2010;103:18–27. doi: 10.1016/j.pbiomolbio.2010.05.005. [DOI] [PubMed] [Google Scholar]
  107. Yang KT, Chang WL, Yang PC, Chien CL, Lai MS, Su MJ, Wu ML. Activation of the transient receptor potential M2 channel and poly(ADP-ribose) polymerase is involved in oxidative stress-induced cardiomyocyte death. Cell Death Differ. 2006;13:1815–1826. doi: 10.1038/sj.cdd.4401813. [DOI] [PubMed] [Google Scholar]
  108. Yang W, Zou J, Xia R, Vaal ML, Seymour VA, Luo J, Beech DJ, Jiang LH. State-dependent inhibition of TRPM2 channel by acidic pH. J Biol Chem. 2010;285:30411–30418. doi: 10.1074/jbc.M110.139774. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Zhang W, Chu X, Tong Q, Cheung JY, Conrad K, Masker K, Miller BA. A novel TRPM2 isoform inhibits calcium influx and susceptibility to cell death. J Biol Chem. 2003;278:16222–16229. doi: 10.1074/jbc.M300298200. [DOI] [PubMed] [Google Scholar]

Articles from The Journal of Physiology are provided here courtesy of The Physiological Society

RESOURCES