Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2008 Aug 1.
Published in final edited form as: Cell Calcium. 2007 May 22;42(2):157–162. doi: 10.1016/j.ceca.2007.03.008

The Long and Arduous Road to CRAC

Monika Vig 1, Jean-Pierre Kinet 1,2
PMCID: PMC2001162  NIHMSID: NIHMS27821  PMID: 17517435

Abstract

Store-operated calcium (SOC) entry is the major route of calcium influx in non-excitable cells, especially immune cells. The best characterized store operated current, ICRAC, is carried by calcium release activated calcium (CRAC) channels. The existence of the phenomenon of store-operated calcium influx was proposed almost two decades ago. However, in spite of rigorous research by many laboratories, the identity of the key molecules participating in the process has remained a mystery. In all these years, multiple different approaches have been adopted by countless researchers to identify the molecular players in this fundamental process. Along the way many crucial discoveries have been made, some of which have been summarized here. The last couple of years have seen significant breakthroughs in the field–identification of STIM1 as the store Ca2+ sensor and CRACM1 (Orai1) as the pore forming subunit of the CRAC channel. The field is now actively engaged in deciphering the gating mechanism of CRAC channels. We summarize here the latest progress in this direction.


Intracellular free Ca2+ is crucial for many cellular processes. Although endoplasmic reticulum (ER) serves to store the intracellular pool of Ca2+, it is limited in its capacity to store and needs to be refilled when depleted. Depletion of ER Ca2+ activates plasma membrane (PM) localized Ca2+ influx channels known as store-operated Ca2+ channels (SOCs). This fundamental act of store-operated calcium influx is believed to have many performers. Nevertheless, calcium release-activated calcium (CRAC) channels have been universally accepted as the protagonists, especially in the context of immune cells. The biggest challenge, as the act unfolds, has been the identity of the key performers, including the CRAC channels themselves. The 20 year old quest began with the hypothesis of store-operated or capacitative calcium entry by Jim Putney in 1986 [1] and subsequent demonstration of calcium release-activated calcium (CRAC) currents (ICRAC) in mast cells by Hoth and Penner in 1992 [2]. Since then many laboratories around the world have been chasing the identity of key players in the act of store-operated calcium influx. Over the years, many genes have been claimed to code for the CRAC channel. Recently, however, two key players have been identified, STIM1, the store Ca2+ sensor and CRACM1 (Orai1), the pore forming CRAC channel subunit. The path to identifying CRAC has been long but not fruitless. Along the way, several other genes with important immunological functions have been identified but the curtains to this act have not yet been drawn. Our laboratory joined the hunt almost 15 years ago and we summarize here multiple different approaches that were adopted by us over these years.

1. The antibody strategy: identification of a new mast cell pathway

Immune cell signaling through antigen receptors, the high affinity receptor for immunoglobulin E (Fc ε RI) and the T cell receptor (TCR), results in the coordinated activation of tyrosine kinases that leads to calcium mobilization from intracellular stores. The resulting activation of CRAC channels localized in the plasma membrane (PM) is crucial for proliferation and cytokine secretion in T lymphocytes [3]. Mast cells require sustained Ca2+ increase for degranulation of allergic mediators and cytokine secretion in response to Fc ε RI aggregation [4]. Thus, one of the very first approaches adopted by us to identify the CRAC modulators involved raising monoclonal antibodies against mast cell surface molecules and selecting clones that inhibited degranulation. This approach resulted in the identification of two PM-expressed molecules CD81 and CD63, that indeed regulate degranulation but they do so not by modulating calcium influx. CD81 and CD63 are important tetraspanin membrane molecules that form multi-molecular complexes with a broad array of molecules including ECM protein-binding beta integrins. Antibodies recognizing CD81 inhibited Fc epsilon RI-mediated mast cell degranulation without affecting tyrosine phosphorylation, calcium mobilization, or leukotriene synthesis [5]. Similar data were obtained with antibodies against CD63. These results revealed an unsuspected calcium-independent parallel pathway of antigen receptor regulation, which was later identified to be Fyn/Gab2/PI3K dependent and which is critical for both degranulation and adhesion. [68]. Both anti-CD63 and anti-CD81 antibodies target the Gab2 pathway to inhibit degranulation while leaving calcium mobilization intact. These findings introduced the possibility of using anti-CD63 and anti-CD81 antibodies as therapeutic agents in allergic disorders.

2. The C. elegans genome strategy: identification of three novel TRPM channels

Having failed to identify CRAC channels with the monoclonal antibody strategy, we designed a bio-informatic approach to screen the entire genome of C. elegans for proteins that were larger than 1000 amino acids. Since many ion channel subunits consist of six transmembrane (TM) helices, this feature was used as an additional criteria to filter the proteins. The resulting candidates were then compared with the EST data base generated from lymphocytes and the overlapping proteins were analyzed for their role in store operated calcium influx.

This approach led to the identification of three members of the transient receptor potential (TRP) superfamily, TRPM2 (previously designated TRPC7 or LTRPC2) [9], TRPM7 (designated as LTRPC7) [10] and TRPM4 [11] with crucial physiological functions.

The members of the transient receptor potential (TRP) superfamily have been in the forefront as likely candidates for store operated calcium entry for over a decade. The gene for TRP channel was first cloned from Drosophila photoreceptors [12], and the TRP proteins were demonstrated to form Ca2+ permeable channels [13], that are activated downstream of G-protein-coupled receptor/PLC signaling pathway [14]. This finding was rapidly followed by cloning and identification of multiple mammalian homologs. The mammalian homologs have been divided into seven subfamilies (TRPC, TRPV, TRPM, TRPN, TRPA, TRPP and TRPML), many of which perform crucial functions in lymphocyte ion metabolism. Many TRP channels are non-selective Ca2+ permeable channels. Since most Ca2+ permeable channels show some sort of feedback regulation by Ca2+ to prevent harmful increases in [Ca2+]i levels, the activity of several TRP proteins, including TRPC3, TRPV5 (ECaC), TRPV6 (CaT1), TRPM2 (LTRPC2) and TRPM7 (LTRPC7), is affected either positively or negatively by changes in [Ca2+]i. It is therefore not surprising that some of the TRP channels have earlier been suspected to constitute the CRAC channel [15]. However, a closer scrutiny in many cases has revealed that these effects are only modulatory. We have briefly described the three TRP channels that we identified and characterized in search of the store-operated channels.

TRPM2: The ADP-ribose gated non-selective channel

TRPM2 is a calcium-permeable channel activated in response to free ADP-ribose (ADPR), a product of NAD hydrolysis and a breakdown product of calcium-mobilizing second messenger cyclic ADP ribose (cADPR) [9]. In addition, TRPM2 has ADPR pyrophosphatase activity, however, this activity is not required for gating of its channel activity [9, 16]. Whether nicotinamide adenine dinucleotide (NAD), a precursor of ADPR, has any effect on channel activation or not is debatable [9, 16]. TRPM2 has also been shown to be activated by micromolar levels of H2O2 and agents that produce reactive oxygen/nitrogen species [17]. Thus a role for TRPM2 in cell death has been proposed. However, these effects are likely to be indirect since oxidative stress can result in the production of ADP-ribose intracellularly [18, 19]. Recently a role for sirtuin, silent information regulator 2 (Sir2), metabolites has been proposed in TRPM2 gating. Sirtuins are protein deacetylases that regulate gene expression, apoptosis, metabolism, and aging [20]. Finally, the two second messengers, cyclic ADP-ribose (cADPR) and nicotinic acid adenine dinucleotide phosphate (NAADP), both of which have been critically implicated in the regulation of calcium signaling in T lymphocytes, strongly activate native TRPM2 channels in Jurkat T cells [2123]. Therefore, it seems plausible that cADPR and NAADP and ADPR, represent physiological co-activators of TRPM2 channels that contribute to Ca2+ influx in T lymphocytes and presumably other cell types that express this channel.

TRPM7: The intracellular Mg2+ sensor

The second channel that we identified using the bio-informatic screen of C. elegans genome is TRPM7, a Ca2+and Mg2+ permeant divalent cation channel. TRPM7 gating appears to be regulated by changes in the intracellular levels of Mg2+ or MgATP, quite similar to native MagNuM or MIC channels [10, 24, 25]. TRPM7’s C-terminal region possesses homology to a unique family of protein kinase domains designated as α-kinase family [2628]. Therefore, in addition to TRPM2, TRPM7 also represents one of the few examples of channels coupled with intrinsic enzyme activity. As in the case of TRPM2’s ADPR pyrophosphatase, the kinase activity of TRMP7 does not appear to influence the channel properties or gating mechanism [10, 24, 25]. On the contrary, the phosphotransferase activity of the TRPM7 α-kinase is modulated by free Mg2+. In fact, it is likely that the proximity of the Mg2+-responsive protein kinase to the channel pore or the site of Mg2+ fluxes across the plasma membrane serves to regulate its kinase activity. Targeted deletion of TRPM7 in DT-40 B cells is lethal, therefore, these channels play a fundamental and non-redundant role in cellular physiology [10].

TRPM4: The calcium activated non-selective (CAN) channel

The third channel we identified, TRPM4, belongs to the calcium activated non-selective (CAN) channel family. These channels are directly activated by elevations in cytoplasmic Ca2+. They support large inward currents carried primarily by Na+ that result in Ca2+ dependent depolarization. Thus, their main function seems to mediate membrane depolarizations to support important cellular responses [2931]. Remarkably, these channels do not undergo voltage or Ca2+ dependent inactivation and therefore are capable of maintaining membrane depolarization. In non-excitable cells, this depolarization decreases the driving force for Ca2+ influx through store-operated Ca2+ channels [11]. This, in turn, can affect the shape and intensity of Ca2+ signals initiated by receptor stimulation. Molecular suppression of TRPM4 converts oscillatory changes of [Ca2+]i into long-lasting sustained elevations in Ca2+ and leads to augmented IL-2 production. Therefore, it is conceivable that physiologically, TRPM4 exerts an important modulatory role on calcium dependent T lymphocyte effector functions such as NFAT translocation and IL-2 secretion [32]. In addition, recently, Vennekens et al. showed that TRPM4−/− bone marrow-derived mast cells had increased Ca2+ influx compared to wildtype cells in response to FcepsilonRI stimulation [33]. As a result TRPM4−/− mast cells showed enhanced degranulation and cytokine secretion. These results further establish a critical physiological role for TRPM4 channels as regulators of Ca2+ influx.

3. The drosophila genome strategy: identification of CRACM1

Since the above described second approach could not identify the CRAC channel, we adopted a third approach that involved performing a completely unbiased whole-genome RNAi screen in drosophila S2 cells. We performed a dye-based assay to measure the real time changes in [Ca2+]i using an automated fluorometric imaging plate reader (FLIPR), in response to a commonly used sarcoplasmic and endoplasmic reticulum calcium adenosine triphosphatase (ATPase) (SERCA) inhibitor, thapsigargin. Analysis of real-time changes in calcium fluxes allowed us to easily cull false positive genes that affected ER release or baseline intracellular calcium levels [34]. 27 genes identified from the dye-based primary high-throughput screen were then re-analyzed in a secondary patch-clamp screen. Patch-clamp analysis identified a single gene olf186-F to be crucial for ICRAC. Around the same time Feske et al. conducted the whole-genome RNAi screen. However, this group designed an assay to study the disruption of nuclear translocation of the nuclear factor of activated T cells (NFAT) in response to thapsigargin mediated elevated [Ca2+]i, using S2 cells stably transfected with NFAT-GFP [35]. Feske et al. combined the results of the high-throughput RNAi screen in Drosophila S2 cells with a modified linkage analysis to identify the mutated gene in two human SCID patients. They also identified olf186-F using the RNAi screen in S2 cells and found the human homolog of this gene to be mutated and located on chromosome 12 in SCID patients [35]. A third report closely followed the above described two reports and confirmed olf186-F to be crucial for SOC entry [36]. The human homologue of olf186-F, FLJ14466 was named as CRACM1 (for CRAC modulator) by us [34] and Orai1 by Feske at al. [35]. CRACM1 is a relatively small TM protein of 32.7 KDa with four predicted transmembrane domains. Over-expression of differentially tagged CRACM1 in HEK cells revealed that both the N- and C-terminal tails are present in the cytosol [34]. Insertion of a tag between the TM3-TM4 loop of CRACM1 confirmed the predicted extracellular location of this loop [37]. Soon after the discovery of CRACM1, a strong functional synergy between STIM1 (the ER Ca2+ sensor described below) and CRACM1 was identified by us [38] and others [36, 39, 40]. So much so that when co-expressed in different cell lines, STIM1 and CRACM1 reconstituted 50–100 fold larger ICRAC currents that were biophysically similar to endogenous ICRAC [38]. These observations pointed towards the possibility that CRACM1 forms the channel itself. Since many ion channels multimerize to form a functional ion pore, we first tested the propensity of CRACM1 to form homopolymers. Co-immunoprecipitation of differentially tagged versions of CRACM1 showed that CRACM1 can form homopolymers with itself [41]. Although multimers of CRACM1 do not form a conventional ring of charged residues, we hypothesized that this may in fact form the basis of the somewhat unconventional pore selectivity of the CRAC channel. Several groups adopted the classical approach to mutate the conserved transmembrane glutamate residues of CRACM1, in positions 106 and 190, to study the changes in the selectivity and permeation of the CRAC channel [37, 41, 42]. The E106Q mutation of CRACM1 not only prevented the reconstitution of large ICRAC currents but also completely suppressed small endogenous CRAC currents. A charge conserving mutation of this glutamate to an aspartate residue E106D (or alanine E106A) resulted in relatively smaller currents and altered selectivity for Ca2+ [41]. The E190Q mutation, on the other hand, resulted in very subtle changes in channel selectivity [41]. These results provided the unequivocal evidence that CRACM1 multimers form the pore-subunit of the CRAC channel with E106 residue conferring the high Ca2+ selectivity. In addition to the two glutamates, there is a string of three aspartate residues in the first extracellular (TM1-TM2) loop. A double mutation in the first two aspartate residues D110/112A increased sodium permeation under calcium free conditions suggesting that these residues may contribute to the selectivity profile of CRAC channels, presumably by facilitating Ca2+ binding to this site thereby contributing to the higher selectivity of Ca2+ ions against monovalent cations [41, 42]. Thus the 20 year long mystery shrouding the store-operated calcium influx finally seems to have been solved by adopting an unbiased approach.

4. STIM1: The store Ca2+ sensor

An important regulatory step in the CRAC signaling pathway is at the level of communication of store depletion to CRAC channels. A couple of years ago, STIM1, an ER resident, single spanning transmembrane protein with a MW of 77 KDa was identified as the crucial calcium sensor coupling the process of store depletion with calcium influx through the CRAC channels [43, 44]. STIM1 possesses an N-terminal EF-hand domain and a SAM domain facing either the lumen of the endoplasmic reticulum or outside when present in the plasma membrane. In addition, a coiled-coil domain and several others are present in the C-term cytosolic tail [45]. Upon store depletion, STIM1 has been shown to relocate to punctae like structures beneath the plasma membrane [43] or in the plasma membrane [46]. Mutating EF hand results in constitutive activation of ICRAC and translocation of STIM1 to punctae or plasma membrane [46, 47]. Whether STIM1 actually inserts into the PM following store depletion has remained controversial. For instance, Zhang et al. showed that the amount of STIM1 pulled down with biotinylated plasma membrane proteins increases upon store depletion [46]. These studies, however, did not rule out the possibility of an indirect pull down of ER STIM1 that is bound to a membrane resident biotinylated protein. Luik et al, on the other hand, showed that STIM1 redistributes into punctae that are located close to PM using TIRF [47]. This study and several others [43, 48] failed to detect the N-terminal tagged STIM1 exposed extracellularly and therefore proposed a role for junctional/punctae aggregated-STIM1 in CRAC channel activation. Recently, Hauser et al. tagged STIM1 with CFP at either the N-term end after the signal peptide (SP) or at the C-terminal tail end and showed that while both the tagged versions of STIM1 translocated upon store depletion, the N-term CFP tagged version did not insert into the plasma membrane whereas C-term tagged STIM1 did, as seen by cell-surface labeling of unpermeabilized cells over-expressing the two different versions [49]. Thus protein tags might affect STIM1 trafficking and prevent its insertion into the plasma membrane, depending on their size and location. Taken together, the above studies suggest that plasma membrane insertion of STIM1 following store depletion is unlikely to be crucial for CRAC channel activation since irrespective of its localization, STIM1 is capable of activating CRAC currents. Interestingly, a fraction of STIM1 is constitutively expressed in the plasma membrane and an earlier independent study assigned a crucial role to PM localized STIM1 in CRAC channel activation by using a monoclonal antibody (mAb) against the N-terminal STIM1 sequence [50]. STIM1 monoclonal antibody suppressed CRAC currents substantially when applied to the outside of intact cells. The same antibody also inhibited the constitutively active CRAC currents in cells expressing the STIM1 EF hand mutant. Contrary to this study, the cytosolic fragment of STIM1, STIM1-CT, was shown to be sufficient to constitutively activate ICRAC in the absence of endogenous PM expressed STIM1 [51, 52]. The role for constitutively expressed STIM1 in the plasma membrane therefore remains to be resolved. Although STIM1-CT was sufficient to constitutively activate CRAC currents [51], the amplification in calcium influx normally seen with the co-expression of WT-STIM1 and CRACM1 was not observed with STIM1-CT and CRACM1 co-expression (our unpublished data). Further studies will likely resolve the role of PM expressed STIM1 in the activation of CRAC currents under physiological conditions.

5. Mechanism of activation of CRAC channels-The next milestone!

Over the years, many general models have been proposed to explain the activation of store operated calcium influx. Although these models predate the discovery of the two crucial molecular players, STIM1 and CRACM1, revisiting some of them in light of the new findings may help in decoding the next big puzzle–the molecular mechanism of activation of CRAC channels. The first model proposed the existence of a diffusible messenger that would link store depletion to CRAC channel activation [53]. Even though the issue of STIM1 and CRACM1 communication currently remains unresolved, a direct interaction between the two molecules would make the diffusible messenger unnecessary. On the other hand, communication over short distances or molecular interaction between ER localized STIM1 and PM localized CRACM1 across two intracellular subcompartments, as proposed by Luik et al. [47] fits better with the secretion-like coupling hypothesis proposed by Gill and colleagues in 1999 [54, 55]. According to this hypothesis, store depletion initiates trafficking of peripheral ER to the PM and when the ER and PM are optimally juxtaposed, coupling between IP3 receptors in the ER (presumably via STIM1) and SOC channels in the PM occurs. A third model proposed that upon store depletion SOCs are inserted into PM via an exocytotic mechanism. This hypothesis was based on the observations that exocytosis blocking drugs inhibited ICRAC [56, 57]. The latest studies have shown that the pore forming subunit of CRAC channel, CRACM1, is constitutively present in the PM [34, 37]. Although these findings rule out the requirement of exocytosis for transporting the CRAC channel in response to store depletion, movement of STIM1 to plasma membrane or junctional ER may require the exocytotic machinery. In fact, a careful analysis of the genes inhibiting calcium influx but not ICRAC (from the recent high-throughput screens) reveals, amongst others, candidates commonly known to be involved in vesicular transport such as SNAP and SNARE (syntaxin5) proteins [34, 36].

Like in the case of many voltage gated channels [58], phosphorylation seems to play a modulatory role in store operated Ca2+ entry [59, 60]. Phorbol ester mediated activation of PKC inhibits store operated Ca2+ entry in the Jurkat cells, human neutrophils, peripheral T-lymphocytes, and rat basophilic leukemia cells [6163]. In contrast to PKC, activation of PKA has shown opposing regulatory effects on CRAC channel function [64]. Serine/threonine phosphatase inhibitors have been reported to inhibit activation of the CRAC pathway [65, 66]. CRACM1 has several putative serine/threonine phosphorylation sites and there is evidence of serine phosphorylation on CRACM1 (our unpublished data). Therefore, it remains to be seen whether phosphorylation plays a role in the activation of CRAC channels.

Acknowledgments

This work was supported by NIH grant 5-R37-GM053950 to JPK and Irvington Institute for Immunological Research award to MV.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • 1.Putney JW., Jr A model for receptor-regulated calcium entry. Cell Calcium. 1986;7:1–12. doi: 10.1016/0143-4160(86)90026-6. [DOI] [PubMed] [Google Scholar]
  • 2.Hoth M, Penner R. Depletion of intracellular calcium stores activates a calcium current in mast cells. Nature. 1992;355:353–356. doi: 10.1038/355353a0. [DOI] [PubMed] [Google Scholar]
  • 3.Lewis RS. Calcium signaling mechanisms in T lymphocytes. Annu Rev Immunol. 2001;19:497–521. doi: 10.1146/annurev.immunol.19.1.497. [DOI] [PubMed] [Google Scholar]
  • 4.Turner H, Kinet JP. Signalling through the high-affinity IgE receptor Fc epsilonRI. Nature. 1999;402:B24–30. doi: 10.1038/35037021. [DOI] [PubMed] [Google Scholar]
  • 5.Fleming TJ, Donnadieu E, Song CH, Laethem FV, Galli SJ, Kinet JP. Negative regulation of Fc epsilon RI-mediated degranulation by CD81. J Exp Med. 1997;186:1307–1314. doi: 10.1084/jem.186.8.1307. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Gu H, Saito K, Klaman LD, Shen J, Fleming T, Wang Y, Pratt JC, Lin G, Lim B, Kinet JP, Neel BG. Essential role for Gab2 in the allergic response. Nature. 2001;412:186–190. doi: 10.1038/35084076. [DOI] [PubMed] [Google Scholar]
  • 7.Parravicini V, Gadina M, Kovarova M, Odom S, Gonzalez-Espinosa C, Furumoto Y, Saitoh S, Samelson LE, O'Shea JJ, Rivera J. Fyn kinase initiates complementary signals required for IgE-dependent mast cell degranulation. Nat Immunol. 2002;3:741–748. doi: 10.1038/ni817. [DOI] [PubMed] [Google Scholar]
  • 8.Kraft S, Fleming T, Billingsley JM, Lin SY, Jouvin MH, Storz P, Kinet JP. Anti-CD63 antibodies suppress IgE-dependent allergic reactions in vitro and in vivo. J Exp Med. 2005;201:385–396. doi: 10.1084/jem.20042085. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.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]
  • 10.Nadler MJ, Hermosura MC, Inabe K, Perraud AL, Zhu Q, Stokes AJ, Kurosaki T, Kinet JP, Penner R, Scharenberg AM, Fleig A. LTRPC7 is a Mg. ATP-regulated divalent cation channel required for cell viability. Nature. 2001;411:590–595. doi: 10.1038/35079092. [DOI] [PubMed] [Google Scholar]
  • 11.Launay P, Fleig A, Perraud AL, Scharenberg AM, Penner R, Kinet JP. TRPM4 is a Ca2+-activated nonselective cation channel mediating cell membrane depolarization. Cell. 2002;109:397–407. doi: 10.1016/s0092-8674(02)00719-5. [DOI] [PubMed] [Google Scholar]
  • 12.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]
  • 13.Hardie RC, Minke B. The trp gene is essential for a light-activated Ca2+ channel in Drosophila photoreceptors. Neuron. 1992;8:643–651. doi: 10.1016/0896-6273(92)90086-s. [DOI] [PubMed] [Google Scholar]
  • 14.Montell C. New light on TRP and TRPL. Mol Pharmacol. 1997;52:755–763. doi: 10.1124/mol.52.5.755. [DOI] [PubMed] [Google Scholar]
  • 15.Yue L, Peng JB, Hediger MA, Clapham DE. CaT1 manifests the pore properties of the calcium-release-activated calcium channel. Nature. 2001;410:705–709. doi: 10.1038/35070596. [DOI] [PubMed] [Google Scholar]
  • 16.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]
  • 17.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]
  • 18.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]
  • 19.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]
  • 20.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]
  • 21.Guse AH, da Silva CP, Berg I, Skapenko AL, Weber K, Heyer P, Hohenegger M, Ashamu GA, Schulze-Koops H, Potter BV, Mayr GW. Regulation of calcium signalling in T lymphocytes by the second messenger cyclic ADP-ribose. Nature. 1999;398:70–73. doi: 10.1038/18024. [DOI] [PubMed] [Google Scholar]
  • 22.Berg I, Potter BV, Mayr GW, Guse AH. Nicotinic acid adenine dinucleotide phosphate (NAADP(+)) is an essential regulator of T-lymphocyte Ca(2+)-signaling. J Cell Biol. 2000;150:581–588. doi: 10.1083/jcb.150.3.581. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.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]
  • 24.Schmitz C, Perraud AL, Johnson CO, Inabe K, Smith MK, Penner R, Kurosaki T, Fleig A, Scharenberg AM. Regulation of vertebrate cellular Mg2+ homeostasis by TRPM7. Cell. 2003;114:191–200. doi: 10.1016/s0092-8674(03)00556-7. [DOI] [PubMed] [Google Scholar]
  • 25.Matsushita M, Kozak JA, Shimizu Y, McLachlin DT, Yamaguchi H, Wei FY, Tomizawa K, Matsui H, Chait BT, Cahalan MD, Nairn AC. Channel function is dissociated from the intrinsic kinase activity and autophosphorylation of TRPM7/ChaK1. J Biol Chem. 2005;280:20793–20803. doi: 10.1074/jbc.M413671200. [DOI] [PubMed] [Google Scholar]
  • 26.Cote GP, Luo X, Murphy MB, Egelhoff TT. Mapping of the novel protein kinase catalytic domain of Dictyostelium myosin II heavy chain kinase A. J Biol Chem. 1997;272:6846–6849. doi: 10.1074/jbc.272.11.6846. [DOI] [PubMed] [Google Scholar]
  • 27.Ryazanov AG, Pavur KS, Dorovkov MV. Alpha-kinases: a new class of protein kinases with a novel catalytic domain. Curr Biol. 1999;9:R43–45. doi: 10.1016/s0960-9822(99)80006-2. [DOI] [PubMed] [Google Scholar]
  • 28.Drennan D, Ryazanov AG. Alpha-kinases: analysis of the family and comparison with conventional protein kinases. Prog Biophys Mol Biol. 2004;85:1–32. doi: 10.1016/S0079-6107(03)00060-9. [DOI] [PubMed] [Google Scholar]
  • 29.Thorn P, Petersen OH. Nonselective cation channels in exocrine gland cells. Exs. 1993;66:185–200. doi: 10.1007/978-3-0348-7327-7_14. [DOI] [PubMed] [Google Scholar]
  • 30.Siemen D. Nonselective cation channels. Exs. 1993;66:3–25. doi: 10.1007/978-3-0348-7327-7_1. [DOI] [PubMed] [Google Scholar]
  • 31.Partridge LD, Muller TH, Swandulla D. Calcium-activated non-selective channels in the nervous system. Brain Res Brain Res Rev. 1994;19:319–325. doi: 10.1016/0165-0173(94)90017-5. [DOI] [PubMed] [Google Scholar]
  • 32.Launay P, Cheng H, Srivatsan S, Penner R, Fleig A, Kinet JP. TRPM4 regulates calcium oscillations after T cell activation. Science. 2004;306:1374–1377. doi: 10.1126/science.1098845. [DOI] [PubMed] [Google Scholar]
  • 33.Vennekens R, Olausson J, Meissner M, Bloch W, Mathar I, Philipp SE, Schmitz F, Weissgerber P, Nilius B, Flockerzi V, Freichel M. Increased IgE-dependent mast cell activation and anaphylactic responses in mice lacking the calcium-activated nonselective cation channel TRPM4. Nat Immunol. 2007;8:312–320. doi: 10.1038/ni1441. [DOI] [PubMed] [Google Scholar]
  • 34.Vig M, Peinelt C, Beck A, Koomoa DL, Rabah D, Koblan-Huberson M, Kraft S, Turner H, Fleig A, Penner R, Kinet JP. CRACM1 is a plasma membrane protein essential for store-operated Ca2+ entry. Science. 2006;312:1220–1223. doi: 10.1126/science.1127883. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Feske S, Gwack Y, Prakriya M, Srikanth S, Puppel SH, Tanasa B, Hogan PG, Lewis RS, Daly M, Rao A. A mutation in Orai1 causes immune deficiency by abrogating CRAC channel function. Nature. 2006;441:179–185. doi: 10.1038/nature04702. [DOI] [PubMed] [Google Scholar]
  • 36.Zhang SL, Yeromin AV, Zhang XH, Yu Y, Safrina O, Penna A, Roos J, Stauderman KA, Cahalan MD. Genome-wide RNAi screen of Ca(2+) influx identifies genes that regulate Ca(2+) release-activated Ca(2+) channel activity. Proc Natl Acad Sci U S A. 2006;103:9357–9362. doi: 10.1073/pnas.0603161103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Prakriya M, Feske S, Gwack Y, Srikanth S, Rao A, Hogan PG. Orai1 is an essential pore subunit of the CRAC channel. Nature. 2006;443:230–233. doi: 10.1038/nature05122. [DOI] [PubMed] [Google Scholar]
  • 38.Peinelt C, Vig M, Koomoa DL, Beck A, Nadler MJ, Koblan-Huberson M, Lis A, Fleig A, Penner R, Kinet JP. Amplification of CRAC current by STIM1 and CRACM1 (Orai1) Nat Cell Biol. 2006;8:771–773. doi: 10.1038/ncb1435. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Mercer JC, Dehaven WI, Smyth JT, Wedel B, Boyles RR, Bird GS, Putney JW., Jr Large store-operated calcium selective currents due to co-expression of Orai1 or Orai2 with the intracellular calcium sensor, Stim1. J Biol Chem. 2006;281:24979–24990. doi: 10.1074/jbc.M604589200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Soboloff J, Spassova MA, Tang XD, Hewavitharana T, Xu W, Gill DL. Orai1 and STIM reconstitute store-operated calcium channel function. J Biol Chem. 2006;281:20661–20665. doi: 10.1074/jbc.C600126200. [DOI] [PubMed] [Google Scholar]
  • 41.Vig M, Beck A, Billingsley JM, Lis A, Parvez S, Peinelt C, Koomoa DL, Soboloff J, Gill DL, Fleig A, Kinet JP, Penner R. CRACM1 Multimers Form the Ion-Selective Pore of the CRAC Channel. Curr Biol. 2006 doi: 10.1016/j.cub.2006.08.085. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Yeromin AV, Zhang SL, Jiang W, Yu Y, Safrina O, Cahalan MD. Molecular identification of the CRAC channel by altered ion selectivity in a mutant of Orai. Nature. 2006;443:226–229. doi: 10.1038/nature05108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Liou J, Kim ML, Heo WD, Jones JT, Myers JW, Ferrell JE, Jr, Meyer T. STIM is a Ca2+ sensor essential for Ca2+-store-depletion-triggered Ca2+ influx. Curr Biol. 2005;15:1235–1241. doi: 10.1016/j.cub.2005.05.055. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Roos J, DiGregorio PJ, Yeromin AV, Ohlsen K, Lioudyno M, Zhang S, Safrina O, Kozak JA, Wagner SL, Cahalan MD, Velicelebi G, Stauderman KA. STIM1, an essential and conserved component of store-operated Ca2+ channel function. J Cell Biol. 2005;169:435–445. doi: 10.1083/jcb.200502019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Putney JW., Jr Capacitative calcium entry: sensing the calcium stores. J Cell Biol. 2005;169:381–382. doi: 10.1083/jcb.200503161. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Zhang SL, Yu Y, Roos J, Kozak JA, Deerinck TJ, Ellisman MH, Stauderman KA, Cahalan MD. STIM1 is a Ca2+ sensor that activates CRAC channels and migrates from the Ca2+ store to the plasma membrane. Nature. 2005;437:902–905. doi: 10.1038/nature04147. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Luik RM, Wu MM, Buchanan J, Lewis RS. The elementary unit of store-operated Ca2+ entry: local activation of CRAC channels by STIM1 at ER-plasma membrane junctions. J Cell Biol. 2006;174:815–825. doi: 10.1083/jcb.200604015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Baba Y, Hayashi K, Fujii Y, Mizushima A, Watarai H, Wakamori M, Numaga T, Mori Y, Iino M, Hikida M, Kurosaki T. Coupling of STIM1 to store-operated Ca2+ entry through its constitutive and inducible movement in the endoplasmic reticulum. Proc Natl Acad Sci U S A. 2006;103:16704–16709. doi: 10.1073/pnas.0608358103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Hauser CT, Tsien RY. A hexahistidine-Zn2+-dye label reveals STIM1 surface exposure. Proc Natl Acad Sci U S A. 2007;104:3693–3697. doi: 10.1073/pnas.0611713104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Spassova MA, Soboloff J, He LP, Xu W, Dziadek MA, Gill DL. STIM1 has a plasma membrane role in the activation of store-operated Ca(2+) channels. Proc Natl Acad Sci U S A. 2006;103:4040–4045. doi: 10.1073/pnas.0510050103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Huang GN, Zeng W, Kim JY, Yuan JP, Han L, Muallem S, Worley PF. STIM1 carboxyl-terminus activates native SOC, I(crac) and TRPC1 channels. Nat Cell Biol. 2006;8:1003–1010. doi: 10.1038/ncb1454. [DOI] [PubMed] [Google Scholar]
  • 52.Lopez JJ, Salido GM, Pariente JA, Rosado JA. Interaction of STIM1 with endogenously expressed human canonical TRP1 upon depletion of intracellular Ca2+ stores. J Biol Chem. 2006;281:28254–28264. doi: 10.1074/jbc.M604272200. [DOI] [PubMed] [Google Scholar]
  • 53.Randriamampita C, Tsien RY. Emptying of intracellular Ca2+ stores releases a novel small messenger that stimulates Ca2+ influx. Nature. 1993;364:809–814. doi: 10.1038/364809a0. [DOI] [PubMed] [Google Scholar]
  • 54.Patterson RL, van Rossum DB, Gill DL. Store-operated Ca2+ entry: evidence for a secretion-like coupling model. Cell. 1999;98:487–499. doi: 10.1016/s0092-8674(00)81977-7. [DOI] [PubMed] [Google Scholar]
  • 55.Rosado JA, Jenner S, Sage SO. A role for the actin cytoskeleton in the initiation and maintenance of store-mediated calcium entry in human platelets. Evidence for conformational coupling. J Biol Chem. 2000;275:7527–7533. doi: 10.1074/jbc.275.11.7527. [DOI] [PubMed] [Google Scholar]
  • 56.Somasundaram B, Norman JC, Mahaut-Smith MP. Primaquine, an inhibitor of vesicular transport, blocks the calcium-release-activated current in rat megakaryocytes. Biochem J. 1995;309( Pt 3):725–729. doi: 10.1042/bj3090725. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.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]
  • 58.Park KS, Mohapatra DP, Misonou H, Trimmer JS. Graded regulation of the Kv2.1 potassium channel by variable phosphorylation. Science. 2006;313:976–979. doi: 10.1126/science.1124254. [DOI] [PubMed] [Google Scholar]
  • 59.Marriott I, Mason MJ. Evidence for a phorbol ester-insensitive phosphorylation step in capacitative calcium entry in rat thymic lymphocytes. J Biol Chem. 1996;271:26732–26738. doi: 10.1074/jbc.271.43.26732. [DOI] [PubMed] [Google Scholar]
  • 60.Zweifach A, Lewis RS. Slow calcium-dependent inactivation of depletion-activated calcium current. Store-dependent and -independent mechanisms. J Biol Chem. 1995;270:14445–14451. doi: 10.1074/jbc.270.24.14445. [DOI] [PubMed] [Google Scholar]
  • 61.Parekh AB, Penner R. Depletion-activated calcium current is inhibited by protein kinase in RBL-2H3 cells. Proc Natl Acad Sci U S A. 1995;92:7907–7911. doi: 10.1073/pnas.92.17.7907. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Montero M, Garcia-Sancho J, Alvarez J. Transient inhibition by chemotactic peptide of a store-operated Ca2+ entry pathway in human neutrophils. J Biol Chem. 1993;268:13055–13061. [PubMed] [Google Scholar]
  • 63.Hardie RC, Peretz A, Suss-Toby E, Rom-Glas A, Bishop SA, Selinger Z, Minke B. Protein kinase C is required for light adaptation in Drosophila photoreceptors. Nature. 1993;363:634–637. doi: 10.1038/363634a0. [DOI] [PubMed] [Google Scholar]
  • 64.Hahn J, Jung W, Kim N, Uhm DY, Chung S. Characterization and regulation of rat microglial Ca(2+) release-activated Ca(2+) (CRAC) channel by protein kinases. Glia. 2000;31:118–124. [PubMed] [Google Scholar]
  • 65.Evans NE, Forth MK, Simpson AK, Mason MJ. Inhibition by calyculin A and okadaic acid of the Ca(2+) release-activated Ca(2+) entry pathway in rat basophilic leukemia cells: evidence for regulation by type 1/2A serine/threonine phosphatase activity. Biochim Biophys Acta. 2005;1718:32–43. doi: 10.1016/j.bbamem.2005.10.005. [DOI] [PubMed] [Google Scholar]
  • 66.Montero M, Garcia-Sancho J, Alvarez J. Phosphorylation down-regulates the store-operated Ca2+ entry pathway of human neutrophils. J Biol Chem. 1994;269:3963–3967. [PubMed] [Google Scholar]

RESOURCES